Hans Christian Ørsted and the Romantic Legacy in Science: Ideas, Disciplines, Practices (Boston Studies in the Philosophy of Science, Vol. 241) - PDF Free Download (2024)

HANS CHRISTIAN ØRSTED AND THE ROMANTIC LEGACY IN SCIENCE

BOSTON STUDIES IN THE PHILOSOPHY OF SCIENCE

Editors ROBERT S. COHEN, Boston University JÜRGEN RENN, Max-Planck-Institute for the History of Science KOSTAS GAVROGLU, University of Athens

Editorial Advisory Board THOMAS F. GLICK, Boston University ADOLF GRÜNBAUM, University of Pittsburgh SYLVAN S. SCHWEBER, Brandeis University JOHN J. STACHEL, Boston University MARX W. WARTOFSKY†, (Editor 1960–1997)

VOLUME 241

HANS CHRISTIAN ØRSTED AND THE ROMANTIC LEGACY IN SCIENCE IDEAS, DISCIPLINES, PRACTICES Edited by

ROBERT M. BRAIN, ROBERT S. COHEN AND OLE KNUDSEN

A C.I.P. Catalogue record for this book is available from the Library of Congress.

ISBN 978-1-4020-2979-0 (HB) ISBN 978-1-4020-2987-5 (e-book)

Published by Springer, P.O. Box 17, 3300 AA Dordrecht, The Netherlands. www.springer.com

Printed on acid-free paper

All Rights Reserved © 2007 Springer No part of this work may be reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, microfilming, recording or otherwise, without written permission from the Publisher, with the exception of any material supplied specifically for the purpose of being entered and executed on a computer system, for exclusive use by the purchaser of the work.

CONTENTS

Foreword

VII

Introduction

ix

List of Contributors

xix

Andrew D. Wilson THE WAY FROM NATURE TO GOD

1

Karen Jelved and Andrew D. Jackson THE OTHER SIDE OF ØRSTED: CIVIL OBEDIENCE

13

Arne Hessenbruch THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

21

Anja Skaar Jacobsen PHRENOLOGY AND DANISH ROMANTICISM

55

Paul Guyer NATURAL ENDS AND THE END OF NATURE

75

Keld Nielsen and Hanne Andersen THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

97

Dan Charly Christensen ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

115

Michael Friedman KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

135

v

vi

CONTENTS

Ernst P. Hamm STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

159

Olaf Breidbach THE CULTURE OF SCIENCE AND EXPERIMENTS IN JENA AROUND 1800

177

Robert Michael Brain THE ROMANTIC EXPERIMENT AS FRAGMENT

217

Lorraine Daston ØRSTED AND THE RATIONAL UNCONSCIOUS

235

Michael Dettelbach ROMANTICISM AND RESISTANCE: HUMBOLDT AND “GERMAN” NATURAL PHILOSOPHY IN NAPOLEONIC FRANCE

247

Trevor H. Levere BETWEEN ENLIGHTENMENT AND ROMANTICISM: THE CASE OF DR. THOMAS BEDDOES

259

Kenneth L. Caneva ØRSTED’S PRESENTATION OF OTHERS’—AND HIS OWN—WORK

273

Roberto de Andrade Martins ØRSTED, RITTER, AND MAGNETOCHEMISTRY

339

Ole Knudsen ØRSTED’S WORK ON THE COMPRESSIBILITY OF LIQUIDS AND GASES, AND HIS DYNAMIC THEORY OF MATTER

387

Frederick Gregory HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE

399

D. M. Knight THE SPIRITUAL IN THE MATERIAL

417

Symposium on H. C. Ørsted and the Romantic Legacy

433

Index

437

FOREWORD

This volume owes its origin to the perception of a puzzling paradox. Hans Christian Ørsted, the great Danish scientist and philosopher, was one of the founders of modern physics through his experimental discovery in 1820 of the interaction of electricity and magnetism—a key step and model for the further unification of the forces of nature. Followers such as Maxwell and Einstein were, and today searchers worldwide are, enchanted by the hope for a completion of that grand program. In addition to Ørsted’s discovery of electromagnetism, his work in science included other fields, chiefly high-pressure physics and acoustics. Moreover, he belonged to that fascinating group of seekers who were deeply engaged in the Romantic tradition of the Nature Philosophers, influenced by Immanuel Kant and by religious, literary, and aesthetic currents. The scientific and philosophical speculations by Ørsted and his circle also quickly stimulated the imagination of other philosophers and scientists. Among the latter were prominently André-Marie Ampère and Michael Faraday, whose work launched the transformation of civilization often called the Second Industrial Revolution, based on the invention of motors, generators, and the pervasiveness of electricity in modern life. But paradoxically, there has long existed, especially in the English-speaking world, what one may justly term an “Ørsted invisibility.” His experiments merit hurried mention in textbooks and class lectures; but few scientists or historians of science have been aware of the important intellectual influences by—and on— Ørsted, of his historical eminence and consequence of his work. Only a select group of scholars have taken up the particular character of thought of the early and mid-19th century that animated Ørsted and his circle, an epoch when researchers in physics, chemistry, biology, and geology enjoyed the pursuit of unity among the sciences as well as with philosophy, religion, the arts, literature, and music. One cause of that “invisibility” was of course that Ørsted published in a variety of European languages, but mostly in Danish and untranslated, which was a barrier to scholarship in the English-speaking community. Therefore, at least two things had to be done. One was to assure that an English translation of a large part of Ørsted’s writings would become available. The book was issued in 1998 by Princeton University Press, Selected Scientific Works of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen, with an introduction by Andrew D. Wilson. With its extensive and authoritative introduction, and the translation of 79 of Ørsted’s articles and other documents, vii

viii

FOREWORD

the book was acclaimed by reviewers and other scholars, presenting to them at last a display of the full panoply of Ørsted’s spectrum of interests and achievements. It was now far more clear than before that Ørsted’s commitment to the principles of Romantic philosophy and aesthetics, acquired through his extensive travels and cosmopolitan connections throughout Europe, remained integral to his scientific research throughout his career. He thus functioned as an illuminating “probe” of European science in the Romantic age. Now the next step could be taken: an international symposium on “H. C. Ørsted and the Romantic Legacy,” held at Harvard University in May 2002. Some two dozen scholars from the USA and six other countries made presentations at the meeting that was cosponsored by the History of Science Department at Harvard University, the Program in Science, Technology, and Society at Massachusetts Institute of Technology, and the Royal Danish Consulate. The organizing committee consisted of Robert M. Brain, Alan M. Brandt, Erwin N. Hiebert, Ole Knudsen, John E. Murdoch, and myself (as chair). The symposium was generously supported financially by an anonymous foundation, as it also had helped in the genesis of the previously mentioned book of the selected scientific works of Ørsted; and both these efforts owed greatly to the key interventions of Ms. Yoyo Tesdorpf Jones. Needless to say, whatever merit this collection has is owing to its authors and its editors, Professors Robert M. Brain, Ole Knudsen, and Robert S. Cohen. After the presentations at the international symposium, the authors undertook the task of revision and completion in light of the discussions at the conference, resulting in the book now before you. We are confident that the volume will serve several ends, especially at this time when the interplay between science and the cultural and social conditions of the times is attracting the attention of scholars as well as lay persons. The collection should deepen our understanding of the relation between the sciences and the philosophical, religious, and aesthetic currents at the time of Ørsted and his contemporaries. Second, it should illuminate the Europewide networks of collaboration and dispute. And it will shed new light on the experimental practice of science during that period. Last not least, one may hope for the next, third step in that project of promoting “visibility”: that this wide-reaching collection will encourage further scholarly work on an era that has too long been largely in the shadows of academe. Gerald Holton Harvard University

INTRODUCTION

As Goethe knew, there is a cosmopolitan bias in science which extends to the writing of its history. The great metropolitan centers of learning dominate the Story, a condition scarcely altered by the turn to local micro-histories. Around 1800, and still to a surprising degree around 2000, geography can be a great handicap in the scientific republic of letters. Still, some work from the scientific periphery proves so luminous, authoritative, and compelling that it gains a place in the metropolitan canon. Such was the case with several of Hans-Christian Ørsted’s scientific works, most notably his discovery of the interaction of electricity and magnetism. This towering pillar of modern physics has earned a place in all histories of science, and set in motion the great researches into electrodynamics and electrotechnology that decisively transformed both physical science and modern life. Yet perhaps because of his origins on the European scientific periphery, the full measure of Ørsted’s scientific achievement, which spanned a broad range of topics in the physical and chemical sciences, remains largely unwritten. Fortunately, this has begun to change, as several fine scholars are involved in probing research into Ørsted’s career and the broader European scientific and philosophical world in which he lived. The present volume aims to further this project, with contributions from many of the leading specialists on Ørsted and the romantic era in the sciences, as well as essays from nonspecialists who have caught some of the fever of this area of research. Ørsted was one of the most prolific and innovative thinkers of his era, the fertile cradling period of European modernity spanning the late Enlightenment and the political upheavals of 1848. Ørsted was, moreover, a European cosmopolitan, a multilingual traveler, and avid correspondent with colleagues across Europe in an age when science still centered largely on the life of the great metropolitan academies. In both his generational and pan-European affiliations Ørsted resembled his contemporary Alexander von Humboldt, who lived from 1769 to 1859. Both men were scientific polymaths and broad-gauge intellectuals, as comfortable in the salon as in the laboratory or academy, with interests ranging widely over literature, the arts, and politics. Not surprisingly, Ørsted and Humboldt befriended and even collaborated with many of the same people, including youthful collaborations with Johann Wilhelm Ritter and the German romantic philosophers of Jena, and with several illustrious French savants of the Paris Academy of ix

x

INTRODUCTION

Science. As scientists, moreover, both men struggled to reconcile the post-Kantian philosophical orientation of the German-speaking world with the experimental and analytical acumen of the French. Yet Ørsted, who mastered the German language in his youth, was more drawn to purely philosophical questions than was Humboldt. Ørsted also remained a more devoted—and skilled—experimentalist than the German savant, whose mature concerns centered more on scientific exploration and methods of gathering, calculating, and representing data from the field. As a result, the tensions between German philosophy and French experiment became more pointed and less clearly resolved with Ørsted than with Humboldt. For this reason many historians have viewed Ørsted as Humboldt’s equal or better as a scientific critic. Both Ørsted’s cosmopolitanism and his independent critical abilities make him an effective probe for the historian of science to discover unseen themes, connections, and faultlines in early 19th-century European science. We have tried to organize both this volume and the conference that engendered it with this feature of Ørsted’s career in mind. We have sought, on the one hand, to account for Ørsted’s formation in his native Denmark, where, apart from his travels, he lived his entire life and exerted great influence over thought, society, and institutions in what historians have called Denmark’s “golden age.” On the other hand, we have endeavored to show how from this “center on the periphery” in Copenhagen Ørsted gained an unusual overview of contemporary European science and by turns encountered frustrating obstacles and decisively shaped the scientific life of the larger metropolitan centers. The organization of the volume follows from these two poles. We open the volume with several essays on Ørsted and the context of his life in golden age Denmark. Most accounts of Ørsted’s scientific and intellectual career begin with his doctoral dissertation on Kant’s philosophy of science, but Andrew D. Wilson’s essay argues that Ørsted’s road to the critical philosophy was laid in his early education in Lutheran catechism and theology. Well before he encountered Kant, and with the salons and academies of Berlin and Paris nowhere in sight, Ørsted received a rigorous schooling both at home and from a German wig maker named Christian Oldenburg and his Danish wife. Wilson shows how Oldenburg imparted a mastery of the German language to Ørsted and his brother Anders Sandøe, along with a thorough grounding in the rationalist Christian apologetics of the German philosopher Christian Wolff and like-minded Danes such as Peder Rosenstand-Goiske and Christian Bastholm. All of these thinkers sought to fend off the criticism and mocking jibes of free-thinkers and skeptics with a theology that reconciled reason and revelation. Wilson demonstrates how this theological movement, particularly Bastholm’s turn to nature and a “theology of Kraft” (force) gave Ørsted a theological perspective which drew him quite naturally to Kant’s dynamical theory and, later, to Schelling’s Naturphilosophie. Breathing so deeply the winds of the Aufklärung, it is hardly any wonder that Ørsted, and other young Danes, devoured the German philosophical and literary works that appeared in the foment of the 1790s. The essays of Arne Hessenbruch, Karen Jelved and Andrew D. Jackson, and Anja Skaar Jacobsen consider Ørsted’s career against the background of the great

INTRODUCTION

xi

changes in Denmark over the course of his lifetime. Like many European countries, Denmark underwent a monumental transformation between 1777 and 1851, moving from an absolute monarchy and broadly manorial economic relations to a constitutional democracy with sweeping political and economic freedoms. Throughout this period, as Jacobsen shows, all of the outstanding European philosophical debates, like those over phrenology and Naturphilosophie, coursed through the academic and intellectual circles of Copenhagen. Hessenbruch situates Ørsted’s career in the middle of the vast changes that swept Denmark in his lifetime. In particular, Hessenbruch shows how Hans Christian and his brother Anders Sandøe, arguably Denmark’s greatest jurist, shaped many of the key legal, social, and institutional conditions of this great transformation. Historians of science have largely ignored the close relations between the Ørsted brothers, described by contemporaries as inseparable as Castor and Pollux. Hessenbruch remedies this lacuna, showing how the fraternal bonds were reflected in the brothers’ shared vision of an enlightened and just society. While they supported the continuation of the absolute monarchy, the Ørsted brothers championed an ethos of personal self-cultivation (Dannelse) and liberal public institutions in both law and education that would mediate communication between the government and the general populace. While Anders Sandøe helped bring about vast political change through the advisory functions of the assemblies of the Estate, Hans Christian worked through the Copenhagen Polytechnic Institute, where he served as a professor and director. The efforts of the Ørsted brothers and their allies provided indispensable conditions for the new Danish institutions that were established after the peaceful abolition of absolutism in 1848—the basic framework that marks the great successes of modern Denmark to this day. Yet in 1848 this was little recognized, as a more radical new generation branded the Ørsteds and their cohort as reactionaries, refusing to acknowledge the liberalizing achievements of the generation of the late Enlightenment. Hessenbruch’s essay shows that there was a good deal more to Ørsted’s life than philosophical pursuits, and suggests many fruitful lines of further historical research for Danish social history, history of technology, and the history of education. No account of the Ørsteds or their cohort can gain any traction without positioning their relation to Kant and to the post-Kantian debates that swept Central European philosophy in their youth. Indeed, the brunt of the slim literature on Ørsted in English concerns the precise nature of his relation to Kant and to postKantian philosophers like Fichte and Schelling. Paul Guyer’s essay takes up some of the key issues of the later Kantian philosophy that informed these debates, namely Kant’s attempt to salvage a role for teleology or purposive explanations in natural and moral philosophy. Kant argued for teleological reason on the basis of heuristic principles which hinged, like much of his later philosophy, on elaborations of the differences between determinative and reflective judgement, and related concepts of constitutive and regulative principles. These questions suffused Ørsted’s dissertation on Kant’s Metaphysical Foundations of Natural Science, which, as Keld Nielsen and Hanne Andersen recount, became the basis of the young man’s philosophy of nature and the basis of his early reputation as a physicist in Denmark. From about 1799 the Ørsted brothers

xii

INTRODUCTION

became the editors of Philosophisk Repertorium, the leading journal disseminating Kantian philosophy in Denmark. Hans-Christian’s interest in the problems of the philosophy of nature of the critical philosophy was unique in his country in those years—most Danish Kantians concerned themselves with the moral and aesthetic which Kant had posed. Dan Charly Christensen traces the key philosophical steps through which Ørsted brought the different kantian strands together by fusing the concept of force with the theory of musical aesthetics. Chladni figures provided the empirical basis for Ørsted’s reasoning, but Kantian categories enabled Ørsted to articulate what would become the constant centerpiece of his philosophical creed throughout his career. While the wider European public knew Ørsted best through his electromagnetic researches, Danes were more likely to associate him with his philosophical acoustics and musical aesthetics, leading Søren Kierkegaard to quip that Ørsted’s face reminded him of “an acoustic figure well-bowed by Nature.” Michael Friedman shows how Kant’s extensive recourse to regulative categories of reason to justify empirical natural science still left much of the natural world unaccounted for—a deficit that post-Kantian philosophers like Schelling, and scientists like Ørsted and Johann Wilhelm Ritter, would strive to fill. Kant’s most egregious shortfall was in chemistry, which the philosopher deemed unscientific. Ørsted, who had trained as an apothecary, sought to carve out a legitimate place for chemistry within the Kantian dynamical theory of matter. The demand became all the more urgent when the new science, electrochemistry, associated with the experiments of Galvani and Volta, became known in Europe during the late 1790s. Kant had not known this work, but for the new generation the new electrochemistry was viewed as nothing short of a revolution in the sciences. Friedman shows how Schelling made intelligent and plausible philosophical moves which aimed to resolve the deep tensions in Kant’s philosophical system and to provide a metaphysical framework for the new kinds of empirical investigation opened up through electrochemistry and electromagnetism. Ørsted and Ritter drew inspiration and encouragement from Schelling’s nature philosophy, even while they found his empirical grasp of the new sciences worthless and even risible. Around 1800 there were other new natural phenomena to reckon with besides the galvanic experiments, and many found Schelling’s philosophy a useful resource for these, too. Once chemistry was provided with a proper philosophical basis it could begin to bolster disciplines such as geology and minerology. The extension of Schellingian matter theory to the earth sciences was precisely the task of Henrik Steffens, Ørsted’s friend and countryman (Steffens was from Norway, which was still under the Danish crown). Steffens journeyed to Germany even before Ørsted, using a Danish stipend intended for study at the renowned Freiberg Mining Academy to undertake a furtive apprenticeship with Schelling in Jena. Although Schelling was the primary focus of his study in Jena, Steffens partook of the rich scientific culture which Olaf Breidbach describes across the university. Naturphilosophie, as Ernst Hamm’s essay shows, gave the young geologist what he needed to recast his science in new terms. Hamm demonstrates many of the steps by which Steffens transformed a vast body of empirical science, with natural phenomena spanning physics, chemistry, and natural history, into a “scientific

INTRODUCTION

xiii

geology,” by which he meant a species of Naturphilosophie, a system of dynamic polarities which narrated both the history of nature and human history unfolding together out of a common inner law. Ørsted appreciated Steffens’ science in his youth (later appraisals were mixed), but pursued his own naturphilosophical quest largely through experimental questions. Both his training as an apothecary and his increasing skill as an experimenter— increased significantly through a stint of collaboration with Ritter in Jena—drew Ørsted to the path of experimentation. But his penchant for experiment also derived from his views on the shortcomings of Schelling’s empirical statements. Everything in Naturphilosophie demanded that the system find its fulfillment in the particular object, either in the material work of art or in the human cognition of material nature. Yet, many of Schelling’s admirers among the romantic circles in Jena stood convinced that philosophy could not fulfill its ambitions alone, without the help of empirical art or science. Robert Brain’s essay contends that Ørsted and other natural philosophers affiliated with Jena romanticism set out to conceptualize experiment as the necessary counterpart to the philosophical system, in direct analogy with the mediating concept of the art work. Ørsted maintained the assumption of a direct analogy between aesthetic intuition and the intuition of nature throughout his life—it was a central theme of his last great statement, The Soul in Nature. Yet just as his formulation of his approach to chemistry was supported by the appearance of galvanic phenomena, Ørsted’s convictions about the primacy of beauty and the sublime in nature were helped by the serendipitous appearance of the acoustic figures of Georg Lichtenberg and Ernst Florens Friedrich Chladni, which seemed a perfect fit for the nature-philosophical insistence on the role of aesthetic intuition in natural knowledge. Ørsted devoted much of his scientific labor to these phenomena, and as Lorraine Daston argues, they formed the basis for his mature notion of a “physics of the beautiful” that would demonstrate a new kind of experimental science based on the cultivated “inner sense” in the scientist. Ørsted, like other romantic nature philosophers, believed that this inner sense was a genuine sense like hearing or sight, with a bodily and mental component. It also straddled the divide between sense and reason, and crucially, Daston argues, conscious and unconscious awareness. Daston shows how Ørsted’s “physics of the beautiful” sought to cultivate this inner aesthetic sense, and in so doing, laid the groundwork for a rational theory of the unconscious that would culminate in Hermann Helmholtz’s musical acoustics of the latter half of the 19th century. Daston’s argument strikes a chord resonant with an observation made by Michael Friedman about the indispensable vanishing act of Naturphilosophie in nineteenth-century science. The crucial concepts and discoveries in both acoustics and electromagnetism that Ørsted bequeathed to the late 19th century grew directly out of Naturphilosophie as their indispensable touchstone, yet this philosophy disappeared almost entirely from the view of those who followed Ørsted (in the case of Helmholtz, it appeared only as an object of scorn and derision). Ørsted’s natural affinity with German thought put him at odds with the very different French approach to natural science. Yet, in an age when ethnonational self-consciousness set the standard for all cultural and intellectual work it was

xiv

INTRODUCTION

not simply enough for a Dane to position himself as an intellectual offshoot of the larger Lutheran nation to the south. Until 1814 Denmark remained an ally of Napoleonic France, and it was against this friendly political backdrop that Ørsted’s first visit to Paris took place. Ørsted was, moreover, deeply impressed by the experimental skill of French scientists, by whose standard he judged experimental science throughout his life. As we mentioned above, Ørsted’s dilemma in many respects paralleled that of Alexander von Humboldt, who was frustrated by French scientists’ lack of a sensibility for nature’s dynamic unity. Michael Dettelbach explores Humboldt’s dilemmas, arguing that the Prussian savant’s own nationalistic impulses were most evident in questions surrounding questions of national religion and the national state, and the place of natural science in both. This was a question that Ørsted also felt keenly, and tried to mediate in Denmark in a variety of contexts, including his dispute with the Lutheran populist theologian N. F. S. Grundvig, and through his labors on behalf of the Copenhagen Polytechnic Institute. Ørsted’s complicated relations with the Parisian pole of the scientific world stood in contrast to his reception in England. One might expect that the relatively robust development of literary and philosophical romanticism in England might have made it fertile soil for Ørsted’s nature-philosophical approach. Trevor Levere’s essay, for example, shows just how involved were British scientists and poets like Thomas Beddoes, Samuel Taylor Coleridge, and Humphry Davy with the German post-kantian and romantic thinkers from whom Ørsted drew inspiration. But Ørsted’s writings did not receive a hearing in Britain until his discovery of electromagnetism, and even then it was through his correspondence and personal connections that his experimental skill gained esteem. Even Michael Faraday confessed to difficulties understanding Ørsted’s electromagnetic experiments in his English-translated articles, although the great natural philosopher eventually mastered and incorporated them into his own researches. Ørsted’s philosophical writings fared even worse, as Gordon McOuat’s oral presentation shows. After years of experimental correspondence with the British electricians and natural philosophers, Ørsted imagined that his philosophical dialogue The Soul in Nature would receive a dear welcome when he sought to have a translation published in 1848. McOuat shows the utter distaste of British natural philosophers for a scientific work that began with beauty and the sublime, then built to discussions of morality, spirit, genius, and reason and religion. The book barely made it into print, and when it did it fell flat. After reading it Charles Darwin noted in his diaries that he found it “dreadful,” a verdict that seemed to speak for all of Britain. Taken together these essays show how Ørsted’s career illuminates the differentiated map of European scientific cultures in the early 19th century, as well as the modalities through which scientific communication crossed or scuppered on national frontiers. But the fact that Ørsted hailed from the periphery should not be taken as a sign that he was a passive participant in these international relations of science. On the contrary, like his countryman Niels Bohr a century later, the Danish savant was a willful and sometimes wily mediator in European science, who tried to turn his peripheral position to his advantage in shaping the course of scientific discussions. Kenneth Caneva’s essay details Ørsted’s strategies with

INTRODUCTION

xv

a wealth of examples, showing how in his correspondence he often tweaked the work of others to serve the position he wished to defend. Reviewing Jakob Joseph Winterl’s Prolusiones, for example, he turned Winterl’s treatise into a species of dynamical Naturphilosophie, which he set up to refute Lavoisier’s chemistry, contrary to all of Winterl’s intentions. Ørsted, as an intrepid traveller, knew an unusually wide variety of scientists’ works through firsthand personal contact, and he showed great skill in shaping the nuances of their presentation to suit his own agenda. Caneva shows how Ørsted managed to use his position to preside over the conventional terminology and usage surrounding different people’s experiments, especially in the decisive translations from the still provincial German to the metropolitan French. Ørsted’s renderings often amounted to a kind of diplomatic compromise, as he expunged what he deemed misuses by Germans of Naturphilosophie in order to make its fundamental philosophical virtues—what he called “its living presence”—more palatable in Paris. Many of the essays in this volume emphasize elements of continuity in Ørsted’s scientific and philosophical creed, while others call attention to important changes in attitude, view, and style over the course of his career. If Ørsted maintained a core commitment to some version of a Kantian or Schellingian philosophy of nature throughout his life, he certainly distanced himself from several features of this worldview with which he flirted in his youth. Roberto de Andrade Martins’ essay shows the close allegiances between Ritter’s researches on the magnetic effects on chemical reactions and the more extravagant features of Schelling’s philosophy, especially the quest for polarities within natural phenomena of all sorts. In Ritter’s zealous investigations the discovery of polarities soon revealed correspondences and symbols or analogies reminiscent of the emblematic worldviews of Renaissance hermeticism. Martins makes manifest Ørsted’s fascination with his friend’s sidereal quest, and highlights the numerous instances when cautious expressions of broad agreement with Ritter’s emblematic polarities entered his own writings. Yet it is clear that Ørsted regarded Ritter’s self-experimentation as excessive, and many of his conclusions hasty and unfounded, an assessment that sharpened in the years following Ritter’s early death. Martins asks whether Ørsted’s changing opinion resulted more from his judgement that Ritter was a poor experimenter or because of changing cultural circ*mstances. The answer might well be both—yet it remains to be determined how much of Ritter’s work the mature Ørsted could still find agreement with. Ole Knudsen’s examination of Ørsted’s work on the compressibility of liquids and gases provides a rare glimpse into the experimental pursuit which dominated the mature scientist’s increasingly rare research time between 1818 and 1845, and which remained the preeminent work in the field until that of P. W. Bridgman in the 20th century. This research showcased Ørsted’s experimental talent and delight in inherent challenges of laboratory work in the absence of any considerations of Naturphilosophie. Nevertheless, as Knudsen shows, Ørsted’s manner of posing the question put to the test the dispute between atomistic and dynamical conceptions of matter—the core issue of physics that had concerned him since his dissertation—which he believed his experiments answered affirmatively in favor of the latter.

xvi

INTRODUCTION

If Ørsted maintained his adherence to the core of his Kantian philosophy of physics, he seems to have held to a similar belief in the spiritual and religious aspirations of Naturphilosophie. Throughout his writings Ørsted observed the many goods that natural science brought both individuals and societies, from social and economic goods to the personal rewards that rigorous intellectual training conferred. But before all these, or rather, subordinate to these, was its theological function, its ability to serve as a “way from nature to God.” If the interest in natural theology was bred in the bone, as we have seen from Wilson’s essay, it fully entered the flesh through the “blessed philosophy of Schelling” and the kindred Naturphilosophen whom Ørsted encountered early in his career and never completely abandoned. Many of Ørsted’s scientific papers, especially those that began as lectures, combine a form of rigorous and modern experimental reportage like that to be found in the Comptes Rendues of the Paris Academy of Sciences with a series of philosophical reasonings that lead one beyond the experiment at hand to a sort of communion with the Godhead. What made Ørsted rare, if not unique, among scientists with a taste for Naturphilosophie was his stern insistence not only on experimental rigor but on sobriety and rationality in philosophical matters. Nobody would dispute that Schelling’s philosophy was rational, indeed rationalist, of course, but the frequent excesses in the writings of both the philosopher and his followers could be easily dismissed as “aesthetic blatherings,” to invoke the words of Emil Du Bois-Reymond. Not so with Ørsted, in whose work Naturphilosophie sits easily with the kind of Naturwissenschaft characteristic of European science after his death in 1851. This presumed virtue might have been a vice for Ørsted’s place in the history of science: too naturwissenschaftlich for the devotees of Schelling, too naturphilosophisch for the practices of the late 19th and 20th centuries, Ørsted nearly vanished in the faultline between them. Frederick Gregory’s essay explores Ørsted’s uniqueness in this regard, pointing out that Ørsted worked in a world where very terms for a scientific researcher in German— Naturhistoriker, Naturforscher, Naturkenner, Naturwissenschaftler—were in transition. Ørsted’s interests in experiment, dynamical concepts, and philosophical systematics put him way out ahead of the fading dominance of natural history around 1800. But when a settlement for these fluctuating categories emerged after 1850 it left Ørsted’s natural philosophy in a strange position, still recognizable in its experimental rigor, but utterly passé in its insistence on philosophical systematics and the spiritual aims of science. As religion, moreover, it was viewed as bland and attenuated, especially compared (in the Danish case) with the fiery popular pietism of Grunvig or the acrid passion of Kierkegaard’s theological polemics. Still, David Knight’s essay reminds us that more British scientists cared about the spiritual around 1850 than we might think (the coupling of “soul and nature” comes from a verse by John Herschel, for example)—they just became increasingly uneasy talking about it in public. Natural theology in Britain was usually more thoroughgoing than Ørsted’s offerings in The Soul in Nature, and it did not obviously share the British versions’ sense of high church holding on for dear life against fire and brimstone, on the one hand, and a growing atheistic radicalism, on the other. Steering clear of those disputes, Knight observes, what scientists like John Tyndall and James Glaisher did retain from Ørsted’s brand of scientific religiosity was the

INTRODUCTION

xvii

vocabulary of the scientific sublime. “Reverence and awe, reason and progress,” Knight writes,” are key words: like a mighty army moves the Church Scientific, but solitariness would be the characteristic of the spiritual experiences of its soldiers.” While organizing the original symposium and in the passage of this book to print we have received help from many people and organizations, whom Ole Knudsen and I would like to gratefully acknowledge. The entire project has been generously supported by an anonymous foundation, and by the History of Science Department of Harvard University, the Program in Science, Technology, and Society of the Massachusetts Institute of Technology, and the Royal Danish Consulate in the USA. We would also like to thank Miss YoYo Tesdorph Jones for essential help and dedication in bringing this project into existence, and Robert S. Cohen for his strong editorial support and sage advice. The Ørsted Symposium Organizing Committee, chaired by Gerald Holton, with Erwin Hiebert, John Murdoch, Ole Knudsen, and Robert Brain, shepherded the project from idea to academic conference and finally to publication. Allan Brandt, as chair of the Harvard History of Science Department, and Peter Galison provided crucial assistance, and Joan Laws, Richard Wright, and Jude Lajoie, helped mightily in myriad ways. We are grateful to Yuri Hospodar whose literary acumen and perseverance improved the manuscript. We would like to especially thank our patient authors and symposium participants, including several who have not contributed chapters to this volume: Frederick Beiser, Christine Blondel, Robert J. Richards, Gordon McOuat, Stuart Strickland, Maria Trumpler, and Albert Gregory. Finally, we would like to offer a unique acknowledgment to Professor Gerald Holton, who conceived the original idea of this project, articulated its vision, and has energetically attended to its every detail from beginning to end. ROBERT MICHAEL BRAIN Harvard University

LIST OF CONTRIBUTORS*

Hanne Andersen, University of Copenhagen, Denmark Robert M. Brain, Harvard University, USA Olaf Breidbach, Universität Jena, Federal Republic of Germany Kenneth L. Caneva, University of North Carolina at Greensboro, USA Dan Charly Christensen, University of Roskilde, Denmark Lorraine J. Daston, Max-Planck-Institut für Wissenschaftsgeschichte, Germany Michael Dettelbach, Boston University, USA Michael Friedman, Stanford University, USA Frederick Gregory, University of Florida, Gainesville, USA Paul Guyer, University of Pennsylvania, USA Ernst P. Hamm, York University, Toronto, Canada Arne Hessenbruch, Massachusetts Institute of Technology, USA Andrew D. Jackson, Niels Bohr Institute, Copenhagen, Denmark Anja Skaar Jacobsen, University of Aarhus, Denmark Karen Jelved, Copenhagen University, Denmark David M. Knight, University of Durham, UK Ole Knudsen, University of Aarhus, Denmark Trevor H. Levere, University of Toronto, Canada Roberto de Andrade Martins, State University of Campinas, Sao Paolo, Brazil Keld Nielsen, Danish Museum of Electricity, Bjerringbrovej, Denmark Andrew D. Wilson, Keene State College, USA

* Institutional identifications as of the time of the Symposium—Ed.

xix

THE WAY FROM NATURE TO GOD The theological foundations of H. C. Ørsted’s philosophy of nature ANDREW D. WILSON

For well over a century, scholars have traced the source of Hans Christian Ørsted’s metaphysics of nature to Immanuel Kant’s dynamical theory of matter and Friedrich Schelling’s romantic Naturphilosophie, both of which he first encountered as a student at the University of Copenhagen during the second half of the 1790s.1 Without question, Ørsted eagerly espoused the general tenets of Kant’s and Schelling’s force-based metaphysics; and, as is well known, Schelling’s Naturmetaphysik was one of the key inspirations behind Ørsted’s discovery of electromagnetism in 1820. To date, however, no one has attempted to explain in any detail why Ørsted so readily embraced an antiatomistic, dynamical physics and metaphysics of nature at a time when the majority of natural scientists were soundly opposed to such an understanding of the physical world (except in the sense that Kantianism was “in the air” in Denmark during the 1790s). In what follows, therefore, I will attempt to offer the beginning of a more specific explanation. I will do so by arguing that Kant’s and Schelling’s Naturphilosophie were consistent extensions in the domain of natural science and speculative philosophy of key theological and religious ideas Ørsted had become familiar with prior to commencing his university studies, and to which he remained committed throughout his long and distinguished career. In this regard, in what follows we shall see that the conceptual foundations of Ørsted’s dynamism and his general philosophy of nature were as much in keeping with a strain of theological thought from the Danish Enlightenment as with Kantian or Romantic metaphysics. In other words, I want to propose that Kant and Schelling provided Ørsted with significant metaphysical elements that fit into a larger, more general Weltanschauung, or total Anskuelse. 1. On 18 January 1850, Ørsted sent a copy of Christliche Dogmatik by Hans Martensen, the Danish Hofpraedikant and Professor of Theology at the University of Copenhagen, to Friedrich Schelling as a small gift to remember himself to his esteemed 1

For an early instance of recognizing the influence of Schelling on Ørsted’s thought see O. Waage, J. P. Mynster og de Philosophiske Bevaegelser paa hans Tid i Danmark (Kbn: C. A. Reitzel, 1867). “…Ørsteds Anskuelse har udviklet sig under Paavirkning af den schellingske Naturphilosophie,” p. 162.

1 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 1–11. © 2007 Springer.

2

A. D. WILSON

friend and intermittent correspondent, whom he had first met 47 years earlier during an extended European tour. The appeal of Martensen’s book to the aging German philosopher was to be found in its sympathetic and somewhat nostalgic treatment of Schelling’s Naturphilosophie. Indeed, Martensen lamented what he described as the “ordinary” system of nature proffered by the new generation of German philosophers, preferring, instead, the “intellectual-poetic, logico-mystical worldview” developed by Schelling and Hegel. Like many intellectuals during the earlier Goethezeit, Martensen also espoused a philosophy of nature in which God was intimately involved in maintaining the order and activity of the physical world. Reminiscent of Schelling, for example, Martensen believed that “God is known in nature as the immanent teleological formal action, as the organizing world-soul (natura naturans). …” Echoing both Schelling and Hegel, Martensen argued that the Divine reveals itself in nature through its “rational action” and the consequent law-determined regularity found, not only in nature, but in history, as well. The presence of divine reason and action in the world led Martensen to conclude that “the universe can no doubt be described as the absolute, insofar as it is a divine fullness, a totality of divine powers (Kraefter) and ideas.”2 In these and other ways, the Christliche Dogmatik was a work wholly consonant with philosophical idealism, wherein, as Bruce Kirmmse has written, “any dualism between reason and nature on the one hand, and revelation on the other” is firmly rejected, and in which, as Martensen himself put it, “God and the world are but two sides of the same unity.” Or, as Ørsted phrased it, “Spirit and nature are one, seen from two different sides.”3Indeed, like other philosophical idealists from his and earlier generations, Ørsted believed that “we should constantly strive to know ever more completely the harmony between reason and revelation.”4 Not surprisingly, then, the views articulated by Martensen, which Ørsted was confident Schelling would read with great favor, also struck a deep resonant chord with him. They were, after all, largely an expression of his own views on the relationship between reason and revelation, and between nature and the divine. The foundations of the natural philosophy and theology Martensen was now presenting had come, in fact, from Ørsted’s lectures on the relationship between nature and spirit he had delivered when Martensen was a student.5 In the letter accompanying Martensen’s Dogmatik, Ørsted told Schelling that he, too, had recently published a book, entitled Spirit in Nature, the purpose of which, he said, was “to prove from empirical science how the collective laws of nature form a rational whole, and how nature itself is a revelation of the creating, living Reason.” Ørsted knew the overarching theme of his book, like the relevant sections in Martensen’s, would appeal to Schelling, since, as Ørsted told him, “The thought is for you as old as your philosophizing.” He then promised that, “as soon 2

3

4 5

Hans L. Martensen, Den Christelige Dogmatik (Kbn: C. A. Reitzel, 1850), pp. 86–87, 80, 99, 85. Ørsted sent Schelling a copy of the German edition. Bruce Kirmmse, Kierkegaard in Golden Age Denmark (Bloomington: Indiana University Press, 1990), p. 176. For the quotation from Martensen, see ibid. p. 80. For the quotation from Ørsted, see To Capitler af det Skjönnes Naturlaere (Kbn: J. H. Schultz, 1845), p. 16. H. C. Ørsted, Imod den store Anklager (Kbn: Andreas Seidelin, 1814), p. 49. H. L. Martensen, Af mit Levnet (Kbn: Gyldendal, 1882), part one, p. 58–60.

THE WAY FROM NATURE TO GOD

3

as a German translation can come out, I will fulfill the pleasurable duty of sending a copy to you, who instructed and inspired me as a young man from afar.”6 The notion that nature is the revelation of a creating, living Reason, was, in fact, the element in Schelling’s Naturphilosophie that both Ørsted and Martensen had found most insightful and noteworthy. In Ørsted’s case, he had spelled this out several decades earlier in a letter dated 1 November 1807 to his brother-in-law, the romantic poet and playwright Adam Oehlenschläger: As is known, Schelling’s merit is to have established Naturphilosophie and with it affected all sciences. The genius of this was not that he constructed nature, and least of all the way he constructed it; but his merit was to view nature as a whole [as] organization. The words: “Nature is a productive product,” would, when properly understood, and they are clearly enough commented on by him, be enough to mark him as a comprehensive and deeply penetrating spirit. He has reiterated this in many different ways. The most beautiful and striking is that nature is nothing other than the revelation of the Divine. You can respond to me that many others have said this before him. I very willingly admit that, and I will say more: All philosophers of great genius have felt this, indeed, said it more or less clearly, but Schelling has articulated it in our time, a time when only a few religious people believed it, but no philosophers perceived it.7

As the passage above suggests, Ørsted was not so enamored with the details of Schelling’s system. He made this even clearer later in the letter, writing, “I have now already said many times that I don’t answer for Schelling’s and Steffens’s constructions, and that there is, again, much to be said against them when I add that I have very often have found them rash.”8 Still, in its broadest outlines, Schelling’s philosophy and his articulation of the connection between God and the world, were much to Ørsted’s liking, and by 1808, he could happily count himself, along with Schelling, among history’s great philosophical geniuses, having publicly presented his belief in the essential link between God, reason, and nature in a dialogue on the natural philosophy of the beautiful, which he published in the transactions of the Scandinavian Literary Society.9 Three years later, in 1811, he again presented his understanding of the relationship between reason, nature, and the divine. Before leaving for a second extended trip abroad, he published a pamphlet, containing a revised and expanded version of the introductory sections to his textbook, Videnskaben om Naturens almindelige Love, the first volume of which had appeared in 1809. In his new introduction, Ørsted inserted the following: 6

7

8 9

Letter from H. C. Ørsted to Friedrich Schelling, dated 18 January 1850. A copy of this letter is in the Ørsted Papirer, folder 2, deposited in the Royal Danish Library. The tenor and content of the essays and dialogues contained in Ørsted’s book were such that it was reviewed in a lengthy essay by J. P. Mynster as a work that could equally well “take its place in the theological literature.” See J. P. Mynster, Blandede Skrivter, 6 Bind (Kbn: Gyldendals Forlag, 1853), vol. 2, p. 218–259. Mathilde Ørsted, ed., Breve fra og til Hans Christian Ørsted, 2 vols. (Kbn: Ch. Linds Forlag, 1870), vol. 1: 230. In 1814 in a polemical booklet directed against N. F. S. Grundtvig, Ørsted again linked Naturphilosophie with helping to revive christian belief, writing, “At Naturphilosophien har bidraget noget til at gjenopvaekke Christendom naegter Grundtvig selv ikke.” See H. C. Ørsted, Imod den store Anklager, (Kbn: Andreas Seidelin, 1814), p. 43. Mathilde Ørsted, Breve, vol.1, p. 231. H. C. Ørsted, “Om Grunden til den Fornøielse, Tonerne frembrage,” in Samlede og efterladte Skrifter, 9 Bind (Kbn: Andr. Fred. Høst, 1852), vol. 3, p. 67–99. During the course of the dialogue, Ørsted’s principal interlocutor twice declares that he considers “Naturen som Aabenbaringen af en uendelig levende og virkende Fornuft.” See pp. 89 and 94.

4

A. D. WILSON All the laws of nature taken together form a unity, which considered in its activity constitutes the essence of the whole world.…If we investigate these laws more closely, we find that they have such complete agreement with reason that we could in truth say that the lawful consistency of nature exists because it is guided by the precepts of reason, or rather, because the laws of nature and the laws of reason are one. The chain of natural laws, which in their activity form every thing’s essence, can thus be considered as a natural thought, or, more correctly, a natural idea. And since all natural laws collectively form a unity, the entire world is an expression of an infinite, universal Idea, which must be one with an infinite, universal living and acting Reason itself. In other words: the world is nothing but the revelation of the Divine’s united Creative Force and Reason. Science must then be studied for its own sake, as our innermost Being’s expression of life, as the comprehension of the Divine.10

The following year, Ørsted published his Ansicht der chemischen Naturgesetze, a major synthetic work which was the culmination of a decade’s worth of work and theorizing. In it, he presented the groundwork for a dynamical chemistry based on Kant’s theory of matter. In a section entitled, “General Considerations concerning the two Grundkräfte,” he again declared that “we see all of nature as the manifestation of an infinite Kraft and an infinite Reason united together, as the revelation of God.”11 Finally, at the end of his life in an unfinished manuscript bearing the title, “The Way from Nature to God,” Ørsted explained that the world “appears to us as the continuous work of a universally penetrating Kraft and Reason, Reason and Kraft; they are in the universe; by them the universe exists and will continue to exist.”12 Indeed, according to Ørsted, Kraft and Reason were the only general constant elements found in nature. As he wrote in the lead dialogue in Spirit in Nature, “We are in agreement that except for nature’s Grundkraefter, the creating Kraefter, there is nothing else constant in things other than natural laws, according to which everything in nature occurs.”13 In other words, according to Ørsted, “the creative Kraft gives a thing its action, Reason gives that action its form.”14 For Ørsted the scientist, therefore, it was by discovering universal natural laws (understood as the revelation of divine Reason) and investigating the actions of universal natural forces (understood as the revelation of divine creative will and creative Kraft), that he believed he could glimpse the thoughts of God, and begin to understand Him through His creation. In this way, according to Ørsted, 10

11

12

13

14

H. C. Ørsted, Første Indledning til den almindelige Naturlaere (Kbn: Johan Frederik Schultz, 1811), pp. 6, 7. Ørsted echoed these thoughts again 44 years later in his To Capitler af det Skjönnes Naturlaere, when he wrote, “Naturlovene i Legemverdenen ere Fornuftlove, Aabenbaringer af een fornuftig Villie.” See H. C. Ørsted, To Capitler, p. 16. H. C. Ørsted, Ansicht der chemischen Naturgesetze (Berlin: Realschulbuchhandlung, 1812) in H. C. Ørsted, Scientific Papers, 3 vols., edited by Kirstine Meyer (Kbn: Andr. Fred. Høst & Søn, 1920), vol. II, p. 157. H. C. Ørsted, “Veien fra Naturen til Gud,” in Samlede og efterladte Skrifter, vol. 3, p. 14. Somewhat earlier in the essay, Ørsted had linked natural laws with divine thoughts and nature with divine revelation: “…og man kunde kalde Naturlovene Naturtanker. Hvad vi kalde Naturtanker er ogsaa Guddomstanker, som vi opdage ved den Aabenbaring, Gud har forelagt os i Naturen.” p. 8. H. C. Ørsted, “Det Aandelige i det Legemlige,” in Aanden in Naturen, 2 Bind, tredie Udgave (Kbn: Andr. Fred. Høst, 1856), vol. I, p. 24. Ørsted had expressed the same view in his Første Indledning of 1811. There he wrote, “Disse Love nu og den Kraft, hvorved de udføres, ere det eneste Uforanderlige i Naturen.” See Første Inledning, p. 4. H. C. Ørsted, “Naturvidenskabens Forhold til adskillige vigtige Religionsgjenstande,” in Aanden i Naturen, vol. 2, p. 42.

THE WAY FROM NATURE TO GOD

5

“All the rules which one can give for the investigation of nature must spring from the fundamental truth: That all of nature is the revelation of an infinite rational Will, and it is the task of science, with finite Kräften, to know as much as possible about it.”15 In this way, too, for Ørsted, the practice of science was, as he told an audience of science students in 1814, a form of religious worship.16

2. The idea that nature is the rational revelation of the Divine and is universally imbued with Kraft by God, while included in Schelling’s metaphysics, did not originate in his philosophy. One need only read Herder’s Gott: Einige Gespräche from 1785 to see this. In Ørsted’s own experience, these ideas had been elements in his religiously oriented education and home life prior to attending the University of Copenhagen. With respect to his religious education, both Ørsted and his brother Anders, who was only one year his junior, recounted how they received their earliest instruction from an old German wigmaker named Christian Oldenburg and his Danish wife. Their schooling consisted of learning to read and write Danish, as well as to read, write, and speak German. They also focused on the Bible, which they heard and read in German, and orally translated into Danish; on Erik Pontoppidan’s edition of Luther’s Kleine Catechismus, which Oldenburg’s wife quizzed them on and most of which they committed to memory; and on pietist religious works, which Herr Oldenburg read to them. As a consequence of their religious instruction, Ørsted and his brother began to write their own sermons and decided, for a time, to become theologians when they grew up.17 The religious orientation of Ørsted’s early education was not restricted to what he received from the Oldenburgs, although it was considerably removed from, if not diametrically opposed to, the Oldenburg’s Pietism. According to Anders Ørsted, “The religious direction in the education which we received in this way in our earliest childhood was furthered by the instruction and example given to us in our parents’ house, as well as by later education following [Ove] Guldberg’s Naturlige og aabenbarede Religion.”18 Part of the instruction the Ørsted children received at home consisted of listening to and reading from the sermons and other works published by Christian Bastholm (1740–1819), the Royal Confessor to the Danish Crown. In regard to Bastholm’s sermons, Anders later recalled how he and Hans Christian “knew and loved his sermons from childhood; every Sunday our mother enjoyed reading one of them, or allowed one of us to read one of them to her.” Anders’s memoirs also relate how he and Hans Christian were thoroughly familiar with other

15

16

17

18

H. C. Ørsted, “Ørsted über das Studium der allgemeinen Naturlehre,” Journal für Chemie und Physik, 6 (1822): p. 475. This is a third, expanded version of the original introduction to Ørsted’s Videnskaben om Naturens almindelige Love (Kbn: Fr. Brummer, 1809). H. C. Ørsted, “Videnskabsdyrkningen, betragtet som Religionsudøvelse,” in Aanden i Naturen, vol. I, p. 137. See H. C. Ørsted, Autobiografi (Kbn: 1828), pp. 1–4; and Anders Sandøe Ørsted, Af mit Livs og min Tids Historie (Kbn: Arne Frost-Hansens Forlag, 1951) for details concerning Ørsted’s early education. A. S. Ørsted, Af mit Livs og min Tids Historie, p. 17.

6

A. D. WILSON

works by Bastholm., and how, when he and Hans Christian arrived in Copenhagen to begin their university studies, they planned to attend Bastholm’s services, only to be disappointed to find that he had stopped preaching publicly by then.19 Through the works of Guldberg and Bastholm, Ørsted was introduced early on to a brand of religious apologetics that had begun to flourish in Denmark during the 1740s in the face of Religionsspottere and Frietaenkeri (mockers of religion and free-thinking). Like the rest of Europe, Denmark had not been immune to the rising influence of “enlightened” attacks on traditional religion and religious faith, which typically pitted critical reason against revelation and scripture in order to expose the superstitious character of the latter. The use of critical reason as the touchstone for truth and understanding at the expense of revelation indeed posed a considerable threat to traditional Christian faith and the authority of the Bible as the revealed word of God. Anders Ørsted, in fact, described the era of his and Hans Christian’s youth as one during which “whoever took part in the Enlightenment…admitted, in general, nothing of value from Christianity other than its ethics, and even believed that one could find the same good and wholesome ethics elsewhere.”20 Evidently, Ørsted’s parents preferred a more balanced approach to the relationship between faith and critical reason, an approach that found its roots, somewhat ironically, in German rationalist philosophy. Earlier in the century, Christian Wolff, the leading disciple of the venerable Leibniz and tutelary representative of philosophical rationalism in Germany, had already begun to try to halt the erosion of Christian religion by attempting to demonstrate the harmonious agreement between faith and reason, a harmony based, however, on the subordination of faith to reason. Danish intellectuals and theologians in search of a justification and foundation for faith drew their inspiration from Wolff. Among those inspired by the Wolffian example were Ludvig Holberg, the dean of Danish letters; and Jens Kraft, professor of mathematics and philosophy at Sorø Academy. Without doubt, however, the most important and influential of the Danish Wolffians was Peder Rosenstand-Goiske, who first introduced Wolff’s philosophy into Denmark in 1742, and who, from 1749 until his death in 1769, was Professor of Theology at the University of Copenhagen. According to Michael Neiiendam, the great scholar of 18th-century Danish religious thought, “an entire epoch built both its faith and its learning” on Rosenstand-Goiske’s continuation of Wolff’s intellectualist program of demonstrating the conformity between reason and revelation.21 In addition to his very popular lectures, Rosenstand-Goiske’s greatest contribution to the apologetic cause was his massive Billige Frie-Tanker over ubilligt Frie-Taenkerie (Reasonable Free Thoughts on Unreasonable Free-Thinking), which he published in weekly installments during 1753 and 1754, and which led steadfast readers through a journey of just over a 1,000 pages. Two followers of Rosenstand-Goiske, who also studied under him at the university were Guldberg and Bastholm, the two religious lights illuminating the Ørsted household. 19 20 21

A. S. Ørsted, Af mit Livs og min Tids Historie, pp. 21–22. A. S. Ørsted, Af mit Livs og min Tids Historie, p. 21. Michael Neiiendam, “Peder Rosenstand-Goiske,” in Dansk biografisk Leksikon, 3rd ed., 16 vols. (Kbn: Gyldendahl, 1979–1984), vol. 12, p. 381.

THE WAY FROM NATURE TO GOD

7

As students of Rosenstand-Goiske, part of the theological mission of both Guldberg and Bastholm was, as Guldberg put it in his Den naturlige Theologie, “to show the intrinsic agreement between reason and Scripture.”22 Their approaches and ultimate intentions, however, were not exactly the same as those of their admired teacher, nor did they themselves follow the same path toward the same goal. In his attempt to reconcile reason and revelation, Guldberg remained quite conservative, being firmly committed to theological supernaturalism, and an absolute belief in the Bible as the direct, revealed word of God. Thus, for him, revelation was synonymous with Holy Scripture. His apologetic approach, in turn, focused on showing how elements of natural religion and theology were consistent with the Divine word. The intrinsic agreement he intended to expose, in the end, was not that revelation had to conform to reason, or even that reason and revelation enjoyed equal standing, but that reason and natural religion were subordinated to faith and revealed religion as embodied in the Bible. Bastholm, on the other hand, differed from his older contemporary, for unlike him, he attempted to place reason and revelation on equal footing. Indeed, from the beginning of his career as a preacher and theologian in the 1770s, Bastholm had set out to demonstrate the parity, mutual dependence, and fundamental unity of reason and revelation. For him, there was no hierarchical order between the two; both found their origin in God and were equally intended by him to demonstrate his greatness. As such, neither was sufficient in itself to accomplish this. He demonstrated this in one of his earliest works, originally written in German in 1770, entitled, Lobrede auf den Messias.23 In this short book, Bastholm combined a traditional, dogmatic explication of the divine nature of Jesus with a critical apparatus of footnotes, which made up about half of the complete text, containing extensive commentary based on nonbiblical historical and mythological sources, and reflecting the ideas and methods of liberal Enlightenment neology. In regard to the Lobrede, Bjørn Kornerup has written that, “In all its heterogeneity, the little text reflects both a fermenting age in which old and new [were] wrestling, and an unsettled personality, who was rootless and homeless.”24 Rather than view Bastholm’s book as a reflection of a theologically disoriented young man, however, it can been seen as an expression of the harmony between biblical revelation and the products of human reason, consistent with the main intentions of the Wolffian apologetic movement in Denmark. As Bastholm wrote in his Den Christelige Religions Hoved-Laerdomme from 1783, a work that is easily recognizable as Wolffian in its organization and contents, “reason and revelation are united to prove [the infinity of God]…. Revelation needs reason, as reason needs revelation.”25 22

23

24

25

Ove Guldberg, Den naturlige Theologie tillige som en Indledning til den Aabenbarede (Sorøe: Jonas Lindgren, 1765), p. 11. Christian Bastholm, Lov-Tale over Messias, oversat af det Tydske ved H. J. Birch (Kbn: F. C. Godiche, 1772). Hal Koch and Bjørn Kornerup, eds., Den Danske Kirkes Historie (Kbn: Gyldendal, 1951), vol. V, p. 365. Christian Bastholm, Den Christelige Religions Hoved-Laerdomme til almindelige Opbyggelse (Kbn: Gyldendals Forlag, 1783), p. 77. With respect to the Wolffian character of this work, see Michael Neiiendam, Christian Bastholm, studier over Oplysningens Teologi og Kirke (Kbn: G. E. C. Gad, 1922), p. 404.

8

A. D. WILSON

Bastholm expressed the same sentiments in his sermons. In them, he wrote, “Is the light of reason perhaps less a gift from the Lord than revelation; or have they not both the same origin; are they not both delivered from the light of the Father. Doesn’t revelation often need the help of reason, just as it again loans reason its light? Revelation requires reason in the same way that reason needs revelation.”26 The reciprocity, dependence, and harmony between reason and revelation is a theme found in many of Bastholm’s works from the 1770s and 1780s, frequently summarized by him in two refrains; namely, “What revelation teaches, reason confirms” and “What reason teaches, revelation confirms.”27 What is most important for us here is that, according to Bastholm, the harmony between reason and revelation was most vividly displayed in nature. In this regard, he argued that, “all of nature bears witness to the infinite wisdom with which it is formed. This witness speaks to our senses, as to our reason.… It is [God], says Jeremiah, who made the world with his power (Kraft), who made the earth with his understanding. If we therefore will only open our eyes to the wonder of God’s great wisdom, and look around us a little, we will soon find the same in nature as Jeremiah did because all of nature calls out to us with a loud voice: God is wise.”28 Moreover, according to Bastholm, if casual observation of nature allows us to begin to discern divine wisdom, then natural science, with its systematic and rational study of God’s creation, further increases our recognition and understanding of God, leading, in turn, to our greater perfection and nearness to Him. In one of the sermons that made up Ørsted’s Sunday instruction, Bastholm described the benefits of natural science, writing, “…the more our sciences expand, the higher our insights ascend, and the more the latter increase, the clearer, more orderly and fundamental our concepts become, the greater is our perfection.…What pleasure doesn’t the lover of truth feel when he views the Creator in creation, when he finds new beauty in it? What excitement doesn’t he feel to discover an unknown wonder, now explaining an obscure law in nature, now bringing new light into a gloomy abyss?”29 Elsewhere, Bastholm declared that, while it is not necessary to be a natural scientist to “recognize and worship” the wisdom of God, it is the natural scientist to whom “it properly belongs to discover the connections of nature and to pursue the goal of making those connections ever clearer.”30 26

27

28 29 30

Christian Bastholm, Aandelige Taler over alle Evangelierne, 2 Bind (Kbn: Gyldendals Forlag, 1779), vol. 2, p. 233. Christian Bastholm, Aandelige Taler, vol. 2, p. 391. Also see Den Christelige Religions Hoved Laerdomme, p. 77. Chr. Bastholm, Den Christelige Religions Hoved-Laerdomme, pp. 51–52. Chr. Bastholm, Aandelige Taler, vol. 1, pp. 635–636. Chr. Bastholm, Den Christelige Religions Hoved-Laerdomme, p. 57. Ørsted expressed much the same view in letters he wrote in February and June of 1808 as he began to work on a long narrative poem on the discovery of balloon flight. As Ørsted described his poetic endeavor in February, he intended to put in the poem, “what natural science has taught me, what inner experiences have developed in my soul, in short, what nature has revealed to me.” He therefore would present the invention of balloon flight as “a natural revelation” (Naturaabenbaring). By June, he had changed his language, shifting his perspective from the natural to the divine. As he told Adam Oehlenschläger, “My purpose is to present this discovery as a divine revelation through nature, which without doubt every discovery is.” For Ørsted, to discover something new in nature is also at the same time to discover something new about God. See Mathilde Ørsted, Breve, vol.1, pp. 261 and 274.

THE WAY FROM NATURE TO GOD

9

Bastholm’s emphasis on the theological and religious importance of observing and studying nature is not surprising. As a student during the 1760s, he had originally devoted himself to natural science and philosophy until his father prevailed upon him to pursue theology and a career in the Danish Church instead. After acceding to his father’s wishes, however, Bastholm never strayed far from his original interests. For him, nature ever remained “religion’s first book.”31 Thoughout most of Bastholm’s career, in fact, nature provided the central context for the apologetic theology he preached and published. Within the context of Bastholm’s natural theology, what is most striking is his understanding and use of the concept of Kraft. Indeed, just as Bastholm viewed nature to be inherently rational, designed according to the plan of divine wisdom, and understood by us through comprehending its laws and realizing how “all things in the world are united and linked together with every other thing; [ i.e. how] the greatest stands in a necessary relation with the smallest,”32 he also understood nature to have been created by divine Kraft and maintained by natural powers that find their origin in God. He first began to articulate what might be descibed as his “theology of Kraft” in his Danish translation of the New Testament, which he published in 1780, and which, like his Lobrede ten years earlier, contained an extensive critical apparatus. Given the Ørsted family’s preference for Bastholm’s writings, this two volume work would have most likely been the edition read by Hans Christian in his youth. In the commentary accompanying the beginning verses of the book of John, Bastholm wrote, “It appears very reasonable to me that he [John] understood the Word to be nothing other than God’s active Will (virksomme Villie) or the infinite Kraft with which he created the world.” He then went on to expound the first verses as follows: “In the beginning…there was an infinite, active, creating Kraft (skabende Kraft) in God; this creating Kraft was God himself. …by this creating Kraft in God all things without exception have their existence.”33 Fourteen years later, he still held to this view, writing, “Every single thing which is produced by the power of nature (Naturens Kraft) according to its orderly way, has no less a ground for its existence in the Divine than the whole of nature.”34 Similarly, according to Bastholm, the Kraft of nature, in turn, finds its ground in God. As he wrote in 1783, “All the Kraft that lies in the creation is the Creator’s and not its own; consequently, the Kraft with which it exists and is maintained, and the Kraft with which it acts necessarily cannot be its own Kraft, but the Creator’s. Without this maintaining Kraft of the Creator, it should be incomprehensible to us how this world’s immeasurable machine should, over so many centuries, be able to exist without weakening or falling apart.”35 The presence and action of this Kraft, in fact, was essential, for, as Bastholm argued, “…it is impossible that the

31 32 33

34 35

Chr. Bastholm, Viisdoms og Lyksaligheds Laere (Kbn: Gyldendals Forlag, 1794), p. 62. Chr. Bastholm, Den Christelige Religions Hoved-Laerdomme, p. 65. Det Nye Testamente oversat efter Grundsproget og oplyst med Anmerkninger af C. Bastholm, 2 Bind (Kbn: Gyldendals Forlag, 1780), vol. 1, pp. 295–296. Chr. Bastholm, Viisdoms og Lyksaligheds Laere, p. 79. Chr. Bastholm, Den Christelige Religions Hoved-Laerdomme, p. 153.

10

A. D. WILSON

world’s matter could be eternal; it must therefore have the ground for its existence in a being outside of it…Reason teaches this…and revelation confirms this.”36 For Bastholm, what linked the existence of matter to its eternal divine ground is the skabende, i.e. creating, Kraft with which God had created and imbued the world, and which continues to create and act according to “the most perfect laws of the highest Wisdom.”37 From this brief overview we can begin to see how Bastholm, through his sermons and other works, presented a natural theology in which nature is understood as the revelation of Divine rational wisdom, and in which the forces that act in nature find their origin in God’s infinite creative Kraft and follow the laws he prescribed for the world according to His purpose. It would not be unfair or inaccurate to say for Bastholm what Ørsted wrote in regard to his own understanding; namely that “the world is nothing but the revelation of the Divine’s united Creative Kraft and Reason,” that “the Creative Kraft gives a thing its action, Reason gives that action its form.” Indeed, both Bastholm and Ørsted saw “all of nature as the manifestation of an infinite Kraft and an infinite Reason united together, as the revelation of God.” I would therefore like to suggest that Ørsted grew up absorbing the natural theology articulated by Bastholm in the sermons he heard and read every Sunday as a youth, as well as in Bastholm’s other works he and his brother read. I would further like to suggest that it was Bastholm’s natural theology and apologetic focus on the inherent connection between reason and revelation that provided Ørsted with a lasting theological foundation from which he pursued his philosophical and scientific interests. I believe it was from the theological perspective provided to him by the works of Bastholm that Ørsted quite naturally was drawn to Kant’s dynamical theory of matter and Schelling’s Naturphilosophie because in them he found a continuation and expansion of Bastholm’s theological apologetics and “theology of Kraft.” Just as religion, for Bastholm, “seeks to direct our desires from the material to the rational, from the earthly to the heavenly, from the visible to the invisible,”38 Kant’s and Schelling’s dynamism in their broad outlines provided Ørsted with a metaphysics that directed him from what he described as the “coarse materialism”39 and atomism of his age, to invisible physical natural forces, and

36

37

38 39

Chr. Bastholm, Den Christelige Religions Hoved-Laerdomme, p. 145. As it turns out, Ove Guldberg had earlier expressed the same view in his Den naturlige Theologie, which Ørsted also studied. For Guldberg, as for Bastholm, “If the world should exist without God, it must be by the world’s own Kraft. But all the Kraft the world has is indeed not its own. It is God’s because He gave it when He created the world.” See Ove Guldberg, Den naturlige Theologie, p. 107. Chr. Bastholm, Aandelige Taler, vol. 2, p. 394. “We think only of the Creator’s omnipotence, but don’t consider that the Being who is infinite in His Kraft is equally infinite in His wisdom, and that this Kraft cannot act without following the most perfect laws of the highest Wisdom.” Chr. Bastholm, Aandelige Taler, vol. 1, p. 334. Mathilde Ørsted, Breve fra of til Hans Christian Ørsted, vol.1, p. 84.

THE WAY FROM NATURE TO GOD

11

from them to metaphysical forces, and ultimately to the infinite creating Urkraft of God.40 As Ørsted wrote to John Herschel in 1849, “I do not deduce anything from metaphysical propositions, but I have tried to show many higher truths as revealed by nature. I would that my readers should take a view through the corporeal world to the mental world.”41 In this regard, the dynamism and metaphysics he encountered during his university years were not fundamental, but were useful in guiding him for the next five and a half decades along the way leading from nature to God, a path he had begun to follow as a boy. Keene State College

40

41

On the concept of Urkraft in Ørsted’s thought see my introductory essay in Selected Scientific Works of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen, with an introduction by Andrew D. Wilson (Princeton: Princeton University Press, 1998), pp. xxxvii–xl. Letter dated 12 October 1849 from Ørsted to Herschel in M. C. Harding, Correspondance de H. C. Örsted avec divers Savants, 2 vols., (Kbn: H. Aschehoug & Co., 1920), vol. 2, p. 405.

Note: Research for this essay was made possible by an NEH Summer Stipend during the summer of 1998.

THE OTHER SIDE OF ØRSTED: CIVIL OBEDIENCE KAREN JELVED AND ANDREW D. JACKSON

Danes refer to the first half of the 19th century as the Golden Age, a time when the interaction between philosophy, art, and science was closer than it is today. The broad spectrum of his interests reveals that Ørsted was very much a man of his time. This is, of course, the primary focus of this symposium. But it was also an era which recognized the fact that scientific techniques and recently acquired scientific knowledge could exert an important and beneficial influence on industry, society, and the lives of ordinary citizens. Ørsted’s activities in this area increased throughout his life, and they are in no small part responsible for Denmark’s continuing affection for him. In this paper we would like to suggest the remarkable scope of Ørsted’s tireless efforts to use the natural sciences in the service of his country. For these purposes, the most important observation regarding Ørsted’s university years is that the breadth of interests and enthusiasms revealed then was central to his later effectiveness as a communicator of science. Indeed, 1798 marked Ørsted’s first attempts in this direction. In a series of four articles Ørsted attempted to introduce contemporary ideas in chemistry to a nonscientific audience. While these articles are far from comprehensive, they remain remarkably readable today.1 They were also well received at the time. After some time spent working as an apothecary in Copenhagen and two years of travel in Germany and France, Ørsted was hired to teach chemistry at the University of Copenhagen. By all indications, he was a popular lecturer. Thus, in 1805 he wrote to the Danish Romantic poet Adam Oehlenschläger2 that “My chemistry lectures have been so well attended this year that there are not enough seats for all. Five or six ladies also attended. You can readily imagine that I did not change my lectures on their account.” In the following year, Ørsted received his first university appointment as Professor Extraordinarius in Physics. In one of his first official acts, Ørsted wrote to university officials requesting a six-month salary advance

1

2

See, for example, Selected Scientific Works of Hans Christian Ørsted, K. Jelved, A. D. Jackson, and O. Knudsen, (Princeton, NJ: Princeton University Press, 1998) for an English language translation of these and Ørsted’s other scientific writings. Ørsted and his brother, Anders Sandøe, became close friends with Oehlenschläger as students in the late 1790s in Copenhagen. Their close relationship was strengthened by A. S. Ørsted’s marriage to Adam’s sister, Sophie.

13 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 13–19. © 2007 Springer.

14

K. JELVED AND A. D. JACKSON

“. . . to meet various necessary expenses occasioned by his new position including improvements to the lecture hall.” At this time, neither chemistry nor physics were represented by separate departments at the University, and Ørsted’s appointment represented the tentative start of an independent physics in Denmark. When not busy teaching, Ørsted was engaged in an exhaustive study of the acoustic figures (Chladni figures) created when a metal plate is stroked with a violin bow and revealed by powder strewn on the plate. While the experiments were relatively familiar, Ørsted’s intentions were new. Consistent with his Kantian convictions, Ørsted wished to demonstrate that the mechanical vibrations of the plates could produce electrical effects. This meticulous and fully professional work was rewarded by membership in the Danish Royal Society (1808) and by the Society’s silver medal. It was also noted in other circles with specific references to acoustic figures appearing in Oehlenschläger’s play Aladdin and later in Søren Kierkegaard’s writing. In 1813 Ørsted returned to Denmark after another trip to Germany. His experiences there had convinced him of the need for major reform of the study of physics in Denmark. His written proposals to the University emphasized the impact that experimental science would have on industry and the population in general. He suggested the immediate founding of a Faculty of Science at Copenhagen University. (Note that this was not done until 1850.) While the study of basic science rather than applied science was and remained Ørsted’s interest, he was emphatic that “theorists are required only for the education of practical experimentalists.” His personal preferences had been stated clearly in his “First introduction to General Physics”3 (1811): We now feel quite vividly how unworthy it would be to make utility the purpose of the study of this or any other science, for when we ask about the usefulness of an object, we thereby reveal that we do not attribute any worth to it in itself but only with regard to something else, which must then be higher. Consequently, if science were to be studied merely for its usefulness, there would have to be something that was more worthy of a rational being than the use of reason or a better part of man than the spiritual, but if this is impossible, then insight is good in itself, and no external justification is needed for wanting to acquire it. Science, then, must be studied for its own sake, as the vital manifestation of our innermost being, as the acknowledgement of the Divine.

In 1814 Ørsted published an article in Latin, Tentamen nomenclaturae chemicae omnibus linguis Scandinavico-Germanicis communis, in which he proposed new chemical terms in Danish, Swedish, German, and Dutch. His aim was to introduce terms that were more immediately recognizable as words of native origin and therefore easier to understand for people with no knowledge of foreign languages. This is why the Danish word for oxygen is ilt, and hydrogen is brint, generated from the words ild (meaning fire) and brand (meaning burn). For confirmation or approval of his newly created words he consulted the eminent Danish linguist, Rasmus Rask, who was less than enthusiastic, but that did not deter Ørsted. Not all the terms he suggested found their way into the Danish language. He also

3

See Jelved et al., Selected Scientific Works of Hans Christian Ørsted …, p. 282.

THE OTHER SIDE OF ØRSTED: CIVIL OBEDIENCE

15

suggested linguistic innovations in other fields and in everyday language—some 2000 words in all. Many were accepted and are still in use. Others did not fare so well. When you look them up in the unabridged Danish dictionary, the only reference is to Ørsted. In spite of Ørsted’s enthusiasm, the general public view of science at the time was not altogether positive. In 1814 the influential Danish clergyman and poet N. F. S. Grundtvig published his World Commentary. He began mildly enough with the categorical assertion that history was the only true science and, warming to his task, continued with the charge that physics had “the denial of God as its first word, sorcery and alchemy its dearest occupations.” Ørsted took up the challenge in an (anonymous) review of World Commentary, and a long, often acrimonious exchange of open letters between Grundtvig and Ørsted followed. The general consensus is that Ørsted emerged victorious. Recognition of Ørsted’s abilities began to accumulate. In 1815 he became the Secretary of the Royal Danish Society and in 1817 he was appointed professor ordinarius. But there was nothing to anticipate the fame that followed his discovery of electromagnetism in 1820. The reception was immediate and enthusiastic. By the first week in September Biot and Arago could report complete verification of Ørsted’s results. On the first Monday in December, Ampère announced his theoretical description of the effect. As a consequence of his lack of mathematical training, Ørsted neither understood nor appreciated Ampère’s contribution. With uncharacteristic sarcasm, he later wrote: “The ingenuity with which this clever French mathematician has gradually changed and developed his theory in such a way that it is consistent with a variety of contradictory facts is very remarkable.”4 Ørsted’s discovery was greeted by immediate acceptance, and there was universal agreement regarding its importance. The years 1822–1823 thus saw Ørsted on a triumphal tour of Germany, France, and England. The visit to England was of particular importance, and Ørsted became aware of the existence of literally hundreds of local societies for the advancement of science. Upon returning home, Ørsted gathered support for the founding in 1824 of a Danish Society for the Advancement of Science (Selskab for Naturlærens Udbredelse (SNU)) to promote the general welfare and to awaken general attention to science.5 The Society was to be (and remains) funded by private subscription. This decision resulted in a profound change of direction for Ørsted’s career. Such public activities began to absorb an increasing share of his time and effort. The goals of SNU were divided. On the one hand, it was dedicated to providing basic science education (i.e. in physics and chemistry). On the other hand, it was concerned with a variety of practical issues including the manufacture of starch, fermentation, brewing, distillation, vinegar production, and tanning. The most visible of SNU’s activities was an ambitious series of public lectures and courses given in Copenhagen and the provinces. During the initial season in Copenhagen, some 200 students

4 5

See Jelved et al. Selected Scientific Works of Hans Christian Ørsted, p. 139. The records from the early years of SNU are to be found at the Niels Bohr Archive in Copenhagen, where they lie largely undisturbed in cardboard boxes—a fascinating source for future research.

16

K. JELVED AND A. D. JACKSON

studied chemistry and physics. At this time Ørsted lectured five evenings each week in addition to his normal University lectures. Lecturers—particularly those outside Copenhagen—were required to spend time each year in Copenhagen to learn of new developments in science and the latest from the industrial world. Ørsted’s instructions to his lecturers were clear enough. They were requested to engage in “a friendly exchange of information and advice rather than one-sided instruction from the scientist.” In addition to its program of public lectures, SNU provided grants to artisans and craftsmen in areas including ceramics, dyeing, and gold- and silver-plating techniques. In spite of the fact that SNU was privately funded, its principal problems were not economic. Ørsted’s greatest difficulty was in finding qualified lecturers. In spite of his demanding public schedule, Ørsted remained active in research. During the first half of the 1820s, he was engaged in a variety of electromagnetic experiments. He also began a long series of experiments aimed at an experimental determination of the compressibility of water and other fluids. The primary experimental obstacle to this determination lies in the fact that the compressibility of water is not markedly different from the compressibility of its container. Ørsted’s ingenious solution to this problem remains useful today.6 His activities as a chemist continued as well, and in 1825 Ørsted succeeded in isolating metallic aluminum. While aluminum generates little excitement today, its discovery was met with excitement. In 1837 an elaborate aluminum parade helmet was fashioned for and presented to King Frederik VII. Ørsted was not equally receptive to all new educational ideas. In 1829, N. F. S. Grundtvig returned from a stay in England with a proposal to found a series of popular high schools to be run locally as “schools for life.” A formal proposal was made to a faculty committee at the University of Copenhagen. Ørsted, serving on this committee, played a central role in its rejection.7 In the same year, G. F. Ursin proposed the formation of a Polytechnic High School in order to provide practical training for artisans. Ørsted was not pleased by the narrow focus of Ursin’s ideas and was again central in the rejection of the proposal. Ørsted, still unhappy about the lack of a natural science faculty at the University, recognized that a Polytechnic High School could provide a suitable alternative. A new proposal for a more academic institution with entrance examinations and greater emphasis on basic science was soon produced, and the Polytechnic High School was quickly created with Ørsted as its first director. As mentioned above, SNU had encountered serious difficulties in coping with the heavy burden of popular education. The Polytechnic High School, with its paid staff of full-time teachers, was now able to assume the bulk of these obligations.

6

7

The fluid to be studied was placed in a vessel fitted with a narrow tube and sealed with a drop of mercury. This vessel was placed in a larger, water-filled container to which pressure could be applied by a piston. The volume of the water could be determined, from the height of the mercury drop, with confidence since its container was subject to equal internal and external pressures. Grundtvig’s high schools were funded from other sources and played a significant role in Danish society. Although they still exist, their importance is now on the wane.

THE OTHER SIDE OF ØRSTED: CIVIL OBEDIENCE

17

Freed of its most time-consuming activities, SNU adopted a slightly different role (both independently and in collaboration with the Polytechnic High School) in bringing science to the public. A remarkably broad spectrum of studies was aimed at technical problems associated with, e.g. the salting of meat and butter, the working of platinum, and the study of building materials. In 1838 new dairy techniques led to the production of some 1,800 kg of one-year-aged cheese. The cheese was auctioned off with great public attention and record prices. SNU also turned its attention to public lectures of a more popular character. There were Sunday lectures for SNU members (and their wives), and there was a regular series of lectures on technological innovations “for the whole family.” Printed summaries of the lectures were always available. Attendance at these events was as high as 2,000 per month, and their contents were frequently reported in the newspapers. The lectures included experimental demonstrations of new devices such as electric motors and steam engines. The 1839 demonstration of the telegraph was reviewed by H. C. Andersen.8 In the same year, the Crown Prince (later King Christian VIII), who had heard the news of Daguerre’s marvelous invention, asked Ørsted to give a public lecture about this new device. Hans Christian Andersen was very enthusiastic: “Our age is the Golden Age of inventions. Oh, would that I, like a Daguerre, could consider how to present mirror images of the heart!”9 Andersen, who was extremely vain, had his portrait taken numerous times. Not so Ørsted, who was more interested in the scientific than the artistic potential of the new invention. SNU produced a variety of publications including a book on Sugar Beets and Beet Sugar (Runkelroer og runkelroesukker) published in 1836. Although it met with initial scorn from Danish farmers, it soon became clear that here was an economically exciting alternative to the dependence on West Indian cane sugar. A profitable production was not far behind. During the 1830s, SNU provided scientific equipment for schools and made the arrangements required to permit school pupils to attend physics and chemistry lectures at the University and the Polytechnic High School. This presence in formal education ended in 1845 when science became an official part of the school curriculum. In 1834 SNU sponsored an exhibition of industrial products, which enjoyed remarkable public success. During the same period, the role of the Polytechnic High School grew dramatically both in its academic impact and its public visibility. These developments were not without a price for Ørsted, who found himself delivering an additional 10 hours of physics lectures per week.10 There were a number of attempts aimed at engaging the Polytechnic High School in the commercial exploitation of its new ideas and inventions. Such attempts were largely unsuccessful due in part to an endless stream of largely administrative problems. The most serious problem, 8

9

10

Ørsted was one of the first to recognize the promise of H. C. Andersen, and their friendship was warm. Andersen’s weekly visits to the Ørsted home continued throughout Ørsted’s life. See Marie-Louise Berner: “Oh, our age is the Golden Age of inventions” in Intersections (Gyldendal, Copenhagen, 2000), pp. 122–137. The intensity of Ørsted’s teaching obligations is remarkable given his scientific accomplishments. It is no less remarkable for being “self-inflicted.” By contrast, Niels Bohr was freed from all University teaching responsibilities in 1927.

18

K. JELVED AND A. D. JACKSON

however, would appear to be the fact that the Polytechnic High School attempted to maintain control of the details of product manufacture rather than spawn “spin-off ” production companies or seek licensing arrangements. At the same time and with greater success, Ørsted sought to create a more academic environment and a more academic treatment of his faculty. The growing popular reputation of the Polytechnic High School can be traced in large part to its active role (through Ørsted) in providing practical scientific advice on a variety of important public issues. Questions regarding the granting of patents and the establishment of monopolies were routinely forwarded to Ørsted.11 Thus, he addressed a variety of specific questions including patents for the telegraph (1846), for asphalt-covered water pipes (1847), and for the creation of a railway line. Without legal or governmental guidance, Ørsted adopted several simple, common sense principles. Patents should be based on scientific priority, and they should cease if new ideas were not exploited by the patent holder. Ørsted demanded (and got) precise descriptions of new products and processes as a prerequisite for patents. He insisted that certain inventions were of such particular value to society (e.g. the telegraph) that patents should belong to the state. These were new and independently formed ideas of considerable value to Danish society. Ørsted was also asked to produce a series of “white papers” regarding technical issues. Not infrequently, these concerned topics on which he had no special expertise. Let us consider one example. In the early 1830s, the bakers of Copenhagen discovered the value of adding 0.1% copper sulphate to bread flour. They were delighted by the lovely crust which resulted. Since bread was sold by weight, they were even more delighted by the fact that the dough absorbed some 10% more water. There was, however, the small matter that copper sulphate is poisonous. The police in Copenhagen thus put a stop to the practice, and Ørsted was asked in 1836 to prepare a new set of rules for the baking of bread. In the introductory notes which accompanied his general directives Ørsted wrote: “If these suggestions are not adopted, it can only be due to the regulations of the bakers’ guild, which render it unnecessary for producers to respond to consumers and which contribute in so many ways to the ignorance of its members.” Ørsted sounds remarkably like a late 20th-century consumer advocate. There were many other activities which kept Ørsted in the public eye. Let me mention only a few. In 1820 he suggested making systematic weather measurements. In 1824 SNU announced a prize for the best proposal for the creation of a “Danish meteorology.”12 In 1827 Ørsted was part of a commission for the establishment of a laboratory for the collection of meteorological data for all of Denmark. Throughout his life Ørsted produced textbooks and popular writing (often of a

11

12

It should be emphasized that Ørsted held no formal government office and had no formal power base. His successful endeavors was due solely to the quality of his arguments and the general conviction that he was providing fair and unbiased advice. Ørsted suggested a new term of Danish-Germanic origin for meteorology vejrligslære, but it never caught on. See Henrik Andersen: “H. C. Ørsted’s Contribution to the Danish Language” in Intersections (Copenhagen: Gyldendal, 2000), pp.138–149.

THE OTHER SIDE OF ØRSTED: CIVIL OBEDIENCE

19

“philosophical” nature). Ørsted even introduced a scale of grades which was used in Danish schools until 1963. A final example of Ørsted’s readiness to oblige when a possibility for popular education presented itself can be found in the Almanac for 1834. This publication reached a wide audience and was the only “literature” to be found in many homes alongside the Bible and the Catechism. Ørsted contributed an article on thunder and lightning, in which he briefly explains what it is and gives advice on how to take precautions against being struck by lightning. In an attempt to change a certain popular distrust of the lightning rod, he tells a story about a church in Siena which was often hit by lightning and therefore equipped with a lightning rod. “The people, who in this place are Catholic and less enlightened than ordinary people in our part of the world, were very upset about this, believing that such a thing would offend God.” God, however, was on the side of progress and saved the church in the next storm, and the good people of Siena were converted to the blessings of the lightning rod. Political correctness is a much later invention. At the end of the same almanac (1834), there is another article by Ørsted. As he explains: “Since there is still some space left, I have been entrusted with the task of writing yet another treatise.” And he goes on to write about “Strong Drink” in a mixture of popular science, moral admonition, and practical advice, including a recipe for punch. He ends the article by saying: “The way in which the juice of our local fruit could best be used to produce wines which are very similar to grape wine would certainly be worth discussing, but I have run out of space.” Ørsted was a scientist and a philosopher. He was also a man, convinced that his science could bring benefits to his country and prepared to invest time and effort to realize this conviction. Indeed, his life was a model of civil obedience. Given Ørsted’s involvement in so many spheres of public life, it is little wonder that literally thousands of Danes joined in a torchlight procession on the occasion of his death in 1851. Copenhagen University/Niels Bohr Institute

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY ARNE HESSENBRUCH

1. INTRODUCTION Hans Christian Ørsted’s life (1777–1851) spanned a tremendous development in science. In 1777 there had been no position for the natural sciences at the University of Copenhagen, by 1851 there was a whole faculty. In Copenhagen in 1777 most people engaging in natural philosophy had had a background in theology, and indeed the educated discourse had been sustained almost entirely by the clergy. By 1851, science related more obviously to manufacturing and industry. In 1777, there had been no general education of the populace and learning was perceived as the kind of training that the young nobility received. By 1851, the more methodical teaching involving exams and graduations were common and opened the door to positions of power and influence. Natural scientists were becoming increasingly professionalized. This development frames Ørsted, his career, and his work. That a natural philosopher be a man of his time seems almost trite, but when he/she is from a peripheral country this is not the case. Outside Denmark, Ørsted is primarily known through those of his writings translated into a major language. Hence, the focus has been on his famous 1820 publication announcing electromagnetism, and to a lesser extent on his 1851 Spirit in Nature.1 Most of the sources that enable the historian to contextualize Ørsted are in Danish only and hence are mostly ignored. Several excellent scholars have made much of the rather limited non-Danish sources at their disposal—some in this book—but I want to argue that something is missed without attention to the actual life that Ørsted lived and the career he pursued. Furthermore, with Ørsted’s ascent on the academic career ladder, he became a most active institutional politician and a prolific writer on the role of science in society. I will argue that this context is highly informative and that it is time for historians of science to give this context at least as much attention as his reading of individual great philosophers. Of course, as a student Ørsted read Kant, Fichte, and Schelling with great enthusiasm, but the enthusiasm

1

Hopefully, the recent translations of Ørsted papers will also help to broaden the view of Ørsted: Selected scientific works of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen; with an introduction by Andrew D. Wilson (Princeton, NJ: Princeton University Press, 1998).

21 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 21–54. © 2007 Springer.

22

A. HESSENBRUCH

should be understood as fueled by the resources that Ørsted recognized in these texts for his own life. And when writing The Spirit in Nature, it is undoubtedly of interest to analyze it as a Kantian text, but it is just as much the swan song of the director of the Copenhagen Polytechnic, outlining his philosophy of education to an uncomprehending and almost hostile world. It is often assumed that Ørsted’s philosophy never changed; for example there are traces of Kantian and Schellingian philosophy in his earliest and his latest writings. My second argument in this paper is that Ørsted did change. His studentdays enthusiasm for German philosophical systems waned as his involvement with public experimentation waxed. His concern with self-control came into focus around 1814, and his interest in “science policy” from the 1820s onward caused him to address the role of science in the emerging civil society. Hans Christian now took on consulting jobs, examining for example national mineral resources, and developing scientific instruments of interest to such enterprises. But most importantly, he became the director of the Polytechnic, the role of which was controversial. This led him to a concern with the roles of science and morality in a good and just society. In this he leaned on his brother, Anders Sandøe, the most prominent Danish jurist of the time, and one anxillary argument to my paper is that the connection between the two brothers matters (a connection that has been completely ignored by scholars of all persuasions, specializations, and language competences). The two brothers remained very close throughout their lives, so close that contemporaries likened them to Castor and Pollux.2 Anders Sandøe never had any children, becoming a widower for the second time in 1824; and he spent much time in Hans Christian’s household. The two of them probably met and conferred most days of the week. Ørsted witnessed dramatic social changes that could not but leave their mark. He was born in an Absolutist state with a minimal and severely constrained public sphere, he died in a democratic state with a growing and increasingly unruly public sphere. Society changed faster than Ørsted was able to. In his youth, he had been a radical, sympathetic to the French Revolution (but not the regicide), and while he did not change a great deal himself, he was outflanked on the left so that in his old age his contemporaries regarded him as a reactionary. In his early career Ørsted had been able to deploy his concept of the sublime in nature as a beacon of godliness, the encounter with which would contribute to the cultivation of all. It was an argument that appealed to the theologically trained academic elite, and

2

C. Hauch, “Hans Christian Ørsteds Levnet,” in Hans Christian Ørsted, Samlede of efterladte Skrifter, Kjøbenhavn : Forlagt af Universitetsboghandler Andr. Fred. Høst : Tryckt hos Kgl. hofbogtrykker Bianco Luno, 1851–1852, 9 vols., vol. 9, pp. 109–183, at p. 150; Krydsfelt—Ånd og natur i Guldalderen, København: Gyldendal, 2000; og Troels G. Jørgensen, Anders Sandøe Ørsted—Juristen og Politikeren, København: Arne Frost-Hansens Forlag, 1952, p. 288. “Brothers connected spiritually as much as by blood,” “Ørsted (Hans Christian),” in Conversations-Lexicon, eller encyclopædisk Haandbog over de i selskabelig Underholdning og ved Læsning forekommende Gjenstande, Navne og Begreber, med Hensyn til Folke- og Menneske-Historie, Politik, Diplomatik, Mythologie, Archæologie, Jordbeskrivelse, Naturkundskab, Fabrik- og Manufacturvæsen, Handel, de skjønne Kunster og Videnskaber, indbefattede tillige de ældre og nyere Tidsbegivenheder, oversat efter den tydske Originals sidste Oplag, med adskillige Forandringer og Tillæg, edited by H. A. Kofod, vol. 28 (Kjøbenhavn: A. Goldin, 1828), pp. 515–541, at 519.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

23

as such helped the advancement of his career, but by the 1840s this argument had worn thin and many clamored instead for a scientific education, especially at the Polytechnic, focused on the skills required for manufacturing. Thus, contextualizing Ørsted requires attention also to the general historical changes, and hence the secondary sources consulted for this paper include general histories, along with institutional histories and biographical material.3 Obviously, my argument that Ørsted was also a Dane and not just a Kantian does not mean that Denmark-Norway-Schleswig-Holstein is the only relevant context, nor that German philosophers do not matter. Ørsted traveled abroad for sustained periods of time, both to Jena, Berlin, Munich, and Paris in 1801–1803, to Berlin and Paris in 1812, and to several places in the German lands, Paris, and London in 1821.4 It is well known that he became very close to Ritter and what later came to be termed romantic German philosophy. His visits to Paris also left their mark; for instance his initiatives for a polytechnic were inspired by the Ecole Polytechnique. And the changes leading from a multilingual absolutist state to a monolingual, nationalistic, and democratic state that framed Ørsted’s Denmark are not strictly local: they mirrored concurrent changes in the German lands and in France. A new civil order came into being in all these areas, all resembling each other. To be sure, there were differences: no revolution ever took place in Copenhagen; the government conceded enough to avoid that. It is usually claimed that the SteinHardenberg reforms in Prussia were a direct response to defeat and occupation by French forces, but similar reforms took place in Denmark. The Copenhagen government did experience defeat in many ways in the early 19th century (e.g. bombardment by the British Navy, loss of the entire fleet, and loss of Norway), but not by the French and without being occupied. When Copenhagen reformers, including the two Ørsted brothers, looked abroad for inspiration, they looked first to the German lands beyond Schleswig-Holstein, and maybe to Sweden, the neighbor and archrival. France, and to a lesser extent Britain, albeit recognized as great powers, were distant and less immediate. But this is merely a caveat; there can be no doubt that the most important context was Copenhagen, where he lived, carved out a career, and defended his worldview against new religious movements, Danish-nationalistic sentiments, and a democratic social order. The structure of this paper is largely chronological. The periods of childhood, youth, and early career are covered in that order, after which the illuminating 1814 clash with the populist preacher, N. F. S. Grundtvig, is analyzed. The chapter on the 1820s and 1830s, when Hans Christian became active in institutional politics and became the iconic public figure of the natural sciences requires a break in the chronology in the form of an excursus. This is because Hans Christian’s contributions to the public debate on the role of science in society leaned on Anders Sandøe Ørsted’s, his jurist brother, elaborations on the requirements for 3

4

I am not the ideal author for this paper. It is not my period, and I am not familiar with the manuscript sources. I have responded to the editor’s request because the historians who are better equipped than I had no interest. I have spent more time than I wanted, but the paper remains sketchy and I look forward to this paper being superseded and improved upon. “Ørsted, Hans Christian,” Dansk Biografisk Leksikon, Sekstende Bind, Woldbye-Aastrup (København: Gyldendal, 1984), pp. 196–202.

24

A. HESSENBRUCH

a new civil order. The excursus covers the career and intellectual development of the latter focusing on what provided a resource for Hans Christian. The last section describes the period of decline, when the social changes went so far that both brothers became generally acknowledged as reactionary. Finally, in the discussion and conclusion, I will pick up the very general issues of trust, professionalism, and cultivation pervading the whole story. It is a theme that has not been addressed for the Danish context, and I will show in very general terms that the development there matches that in Britain.

2. THE ØRSTED BROTHERS’ EDUCATION, 1777–C.1800 2.1. Childhood Hans Christian and Anders Sandøe’s father was the apothecary of a market town, Rudkøbing, on the island of Langeland. Rudkøbing is 80 km (50 miles) from German-speaking Kiel, as the crow flies, and almost twice that from Copenhagen. The Rudkøbing harbor will have ensured that the boys encountered ships and sailors from elsewhere, in particular also from German-speaking parts of Schleswig-Holstein, then still duchies of the Danish-Norwegian King. They heard German spoken on a daily basis. Up until the second half of the 18th century the structure of agricultural life had barely changed, and some 80% of the population was rural (the degree of urbanization was no lower than elsewhere in Western Europe). The landed gentry owned most of the land. They also owned the copyhold farms and agricultural tools. Copyholders paid for the use of land, houses, and tools with agricultural products, money, and labor. They were not to leave without permission and they had to provide soldiers in times of war. In the second half of the 18th century the population increased and for the first time ever was not kept in check by disease or famine. Consistently high prices for agricultural products combined with population growth led to dramatic reforms, especially in the 1780s. Gradually more and more land was privatized and enclosed, the village community broken up, and farms moved out on to their own plots. The symbolically important freedom of movement of the peasant (abolition of adscription) was decreed in 1788 and took place over the next 12 years; the slave trade was made illegal in 1792 and more and more obligations in the form of labor were transformed into duties in the form of money.5 By and large, this matches the reforms elsewhere, for example those under Joseph II.6 The Ørsted brothers must have experienced firsthand the increasing enclosure, privatization, and improvement of agriculture that turned rural life upside down prompting considerable resistance and leaving some who had found shelter in the village community to fend for themselves. The modernization of agriculture 5

6

Ole Feldbæk, Den lange fred—1700–1800—Gyldendal og Politikens Danmarkshistorie, Bind 9 (Copenhagen: Nordisk Forlag & Politikens Forlag, 1990). Carlile Aylmer Macartney, The Habsburg Empire, 1790–1918, London: Weidenfeld & Nicolson, 1971; Hajo Holborn, A History of Modern Germany (Princeton, NJ: Princeton University Press, 1982).

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

25

led also to vastly increased agricultural production and greater affluence—which impressed the brothers; as Hans Christian was later to put it: “free hands give rich yield.”7 As students, the brothers were taken by the ideals of egalitarianism while observing many clear class distinctions around them. The abolition of adscription and slave trade also affected them.8 Taxation ceased to be farmed out, instead becoming a task of a civil service, with increasingly professional surveyors measuring land as precisely as possible. Indirect taxation, levies raised wherever a bottleneck was found in the traffic of goods (e.g. town gates, harbors, mills, turnpikes), also grew markedly in this period.9 This expanded role of the state will have been visible in the boys’ everyday life. The budding natural philosopher and jurist must have seen measurements performed with increasing attention to precision and prescribed in evermore regulatory detail in order that the levy be seen as fair. That measurement for taxation purposes was an issue of morality (and for example discussed repeatedly in the Bible) was a commonplace.10 The brothers received no formal schooling as there was none. Not until 1814 did primary education become compulsory from the age of 7 until confirmation (the Lutheran ceremony at approximately age 14 confirming the faith declared by the parents at baptism).11 Mostly they studied with a German wig maker and his wife, and assorted others, including a local minister and the local judge. The wig maker’s wife taught them the catechism, and the wig maker himself taught them German, addition, and subtraction. From others they learnt multiplication and division, and also some Latin, French, and English. The atmosphere was most devout and the brothers aspired to become theologians, practising the writing and giving of sermons.12 The wig maker again and again emphasized German cultural superiority. Politeness was taught, including polite behavior toward those on a lower social rung.13 The kind of cultivated behavior the two brothers learnt to appreciate can be gauged by Hans Christian’s 1802 description (at the age of only 25) of a dinner party not to his liking: Nothing is more execrable to me than wicked songs and a lack of higher-spirited entertainment. A carouse is loathsome if not a true solemnity shared with educated human beings, where we drink moderately and achieve a higher turn whereby my soul is opened to more joys and also to higher feelings of every kind. First cheerful ditties and then more serious, such as Schiller’s Ode to Joy; that is what I wish.14 7 8

9

10 11

12 13 14

From the poem: “Fædrelandssang,” Samlede og efterladte Skrifter, vol. 4, 28. Anders Sandøe Ørsted, Af mit Livs og min Tids Historie—Udgivet med Forkortelser i Hundredaaret for Oprettelsen af det Anders Sandøe Ørstedske Prismedaille Legat ved Legatets Bestyrelse (København: Arne Frost-Hansens Forlag, 1951), pp. 18–19. Arne Hessenbruch, “The spread of precision measurement in Scandinavia 1660–1800,” in The Sciences in the European Periphery During the Enlightenment, edited by Kostas Gavroglu (Dordrecht The Netherlands: Kluwer, 1999), pp. 179–224. Witold Kula, Measures and Men (Princeton, NJ: Princeton University Press, 1986). Claus Bjørn, Gyldendal og Politikens Danmarkshistorie, Bind 10: Fra reaktion til grundlov (Copenhagen: Gyldendal, 1990), p. 49. “Ørsted (Hans Christian),” in Conversations-Lexicon (n. 2), at 517. C. Hauch, “Hans Christian Ørsteds Levnet” (n. 2), at 112. Mathilde Ørsted (ed.), Breve fra og til Hans Christian Ørsted—Første Samling (Kiøbenhavn: Th. Linds Forlag, 1870), p. 39.

26

A. HESSENBRUCH

The two brothers clearly ended up with much knowledge by the standards of the day, and with the skills of civility and foreign languages to boot. In the 1790s, the frustration by Danes to be treated as culturally and socially second-rate by the aristocratic, German-cultured political elite began to emerge.15 As the historian Ole Feldbæk put it: “The frustration may well have been increased by the obviously different cultural levels of the Danish clubs and the literary salons in the mansions of the Bernstorff, Schimmelmann, and Reventlow families.”16 The two Ørsted brothers will not have experienced that frustration, since they were well placed to move in those circles, and Hans Christian was later to develop an excellent rapport with Count Schimmelmann in particular. 2.2. Youth In 1793 they moved to Copenhagen to prepare for and take a qualifying exam for the University. In 1797 Hans Christian graduated in pharmacy and in 1799 he took a doctorate in natural philosophy, the same year that Anders Sandøe received his law degree. Both excelled. Hans Christian won gold medals in the essay competitions in aesthetics in 1796 and in medicine in 1797. Anders Sandøe’s essay in response to the 1798 competition on the relation between ethics and law likewise won a gold medal.17 At the time, the theological department dominated life at Copenhagen University. Lutheranism was the only sanctioned religion in the kingdom, apart from a small number of Jewish communities and members of the Reformed Church. Lutheranism was interwoven with the state, and the King presided over it with all the symbolism attending royal and absolute power. The maybe 1,000 clergy constituted the majority of the learned world, and much of the state administration was in their hands, including registration of births, marriages, and deaths.18 Public debate had begun in a very limited way in the 1740s. The audience was small, journals economically weak, amateurishly edited, and provincial in tone. The debate grew steadily during the second half of the 18th century to include topics such as the place of the individual and of groups in society and in relation to the state, the duty of Absolutism to listen and be advised, and the relevance of being one people with a shared language and a shared past.19 The Absolutist state

15 16 17

18 19

This is a very early manifestation of proto-nationalism, maybe the first. Ole Feldbæk, Den lange fred—1700–1800 (n. 5), pp. 335–336. “Ørsted (Anders Sandøe), in Conversations-Lexicon, eller encyclopædisk Haandbog over de i selskabelig Underholdning og ved Læsning forekommende Gjenstande, Navne og Begreber, med Hensyn til Folke- og Menneske-Historie, Politik, Diplomatik, Mythologie, Archæologie, Jordbeskrivelse, Naturkundskab, Fabrik- og Manufacturvæsen, Handel, de skjønne Kunster og Videnskaber, indbefattede tillige de ældre og nyere Tidsbegivenheder, oversat efter den tydske Originals sidste Oplag, med adskillige Forandringer og Tillæg, edited by H. A. Kofod, vol. 28 (Kjøbenhavn: A. Goldin, 1828), pp. 541–563, 546. Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 155. Ole Feldbæk, Den lange fred (n. 2), pp. 209–210. There was also a genre of useful knowledge, perhaps best compared to current do-it-yourself/home improvement literature, Michael F. Wagner, Det polytekniske gennembrud—Romantikkens teknologiske konstruktion 1780–1850 (Århus: Aarhus Universitetsforlag, 1999), p. 106.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

27

kept close tabs on the development, and while it did explicitly censor, it also had an interest in allowing a certain measure of public debate so as to know what was going on. In addition to the formal censorship, self-censorship took place: prior to the existence of an affluent public capable of purchasing critical journals, the authors were usually dependent upon a position within the civil service.20 The nature of state control of public discourse at the time of the Ørsted brothers’ childhood may be seen from the language in a 1780 regulation that the Head of Police was to bring to presidents of the many newly founded clubs in the capital: The so-called clubs that have been introduced in this city have nothing reprehensible in and of themselves. But the youth might well fall into idleness, disregard of their high duties, and profligacy with their for the state expensive time. Such gatherings might even lead to more dangerous excesses…. Presidents ought to be civil servants or at least men of respected age and a settled nature with awareness of decent behavior. [The clubs should close no later than 11pm and the young men], this so lubricious age group, should be led by reasonable and respected company to virtue, wisdom, politeness, and thrift.21

In the second half of the 18th century, the discourse on Reason had put the Danish state church on the defensive. Put on the defensive by philological and historical critiques of the Bible, and especially of its tales of omens and miracles, Church representatives compromised. The compromises involved abandoning evermore details of the Holy Writ. By the 1790s, in the wake of the French Revolution, ridicule of the Church and its rituals became a part of debates in Copenhagen, until the state clamped down hard. For a while the brothers accepted the widespread opinion among the “enlightened class” in Copenhagen that only the ethics of Christianity has any justification, and that one might really do away with religion altogether.22 Quite possibly the decapitation of the French king and the widespread abhorrence this event prompted in their environment prompted their return to a belief in revealed Christianity. In subsequent years the civil service grew, creating much new delegation of power. For example, market towns had representatives with a say in local decisions, and from 1797 onward the burghers (some 10% of the population) elected these representatives. In the same years, many new commissions were created, e.g. harbor commissions elected locally. Burghers could become Post Master, City Treasurer, or be given command over civil regiments.23 The Ørsted brothers graduated from university just as people with their kind of background gained such opportunities. At the same time, the status of education went through a transformation: In the 1790s the aristocratic and cosmopolitan form of education no longer enjoyed a monopoly. The academic education, marked by exams, increasingly rivaled its status. In the 1800s and 1810s, academics increasingly claimed authority for the determination of social development.24 20 21 22

23 24

Ole Feldbæk, Den lange fred (n. 2), p. 318. Ibid. p. 321. Anders Sandøe Ørsted, Af mit Livs og min Tids Historie—Udgivet med Forkortelser i Hundredaaret for Oprettelsen af det Anders Sandøe Ørstedske Prismedaille Legat ved Legatets Bestyrelse (København: Arne Frost-Hansens Forlag, 1951), p. 21. Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 68. Ibid. p. 163.

28

A. HESSENBRUCH

This does not amount to egalitarianism. All women and 90% of the town’s population had no share in the new opportunities. Furthermore, the law still regarded the peasantry as distinct from the rest of the population, a state of affairs connected with the direct taxation of the land. As a result, legislation, courts of law, and the administration all treated the peasant as in need of tutelage. One might talk about the emancipation of a narrow band of professionals, who saw the opportunities coming to them on the basis of their skills or education, not God-given by class or birth. The Ørsteds did not conceive of themselves as inherently different from the bulk of the population, a radical position compared with the de facto legal distinction between peasants and others. But they did see themselves as superior qua educated and cultivated human beings. This was exactly how the majority of the late 18th-century clergy saw its role, and it was common to stress the role of the pastor as his congregation’s teacher, a good pastor both as capable clergyman and secular civil servant, an ideal in terms of pater familias and agriculturalist. Many pastors in this generation labored hard to build up an infrastructure for the poor, for schools, and for adult education. Many introduced the new kinds of agriculture that enclosures facilitated: field rotation, ploughing, and rejuvenation of the soil.25 This 1,000–1,500-strong clergy dominated the public sphere and constituted the Ørsted brothers’ role models. The brothers summoned considerable youthful enthusiasm for German philosophers, such as Kant, Schelling, and Fichte, because they saw the relevance of the philosophy to the Copenhagen around them. Anders Sandøe (and probably Hans Christian too) first encountered Kantian philosophy in the journal Repertorium for Fædrelandets Religionslærere (Repertory for the religion teachers of the fatherland), in which the Kantian philosophy was celebrated, and then from Schelling’s lectures in 1796. We were deeply affected by the Kantian moral principle, and the associated teaching on duty as the only driving force for human action of any value. All my fellow students and I never doubted the truth of the scientific and especially ethical qualities of Kant’s system. We considered that he had developed the nature of freedom satisfactorily, based on the conviction that the human soul is free in the real and full sense of the word, and that therefore humans are undoubtedly responsible for their own actions.26

In other words, Kant’s system fit their political convictions that social hierarchies should be built on merit, not birth. Given the proper education, the general populace could be elevated. Social hierarchies are not fixed or defined by birth, and in principle all humans are equal. A mark of this equality is that all humans are responsible for their own actions, unlike animals. By 1799, the brothers had finished their University education and had to find a way to make a living. Philosophy might do well for dinner parties and for getting published, but it did not help directly to make ends meet. From 1800–1801 Hans Christian got an academic foothold teaching pharmacy at the University and the Academy of Surgeons (without a salary) while running the apothecary shop in

25

26

Ibid. p. 155–157; Dan Ch. Christensen, Det moderne projekt: teknik og kultur i Danmark-Norge, 1750–(1814)–1850 (København: Gyldendal, 1996). Anders Sandøe Ørsted, Af mit Livs og min Tids Historie (n. 22), p. 25.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

29

the owner’s absence. Anders Sandøe applied in vain for a teaching position in the Faculty of Law. In 1801 he accepted a position as a judge in the city court of Copenhagen. Slowly, Hans Christian also got his finances on more of an even keel. He received a travel stipend to do the expected European tour, spending 1801–1803 in Jena, Berlin, Munich, and Paris. Upon his return he garnered a position teaching physics, and in 1806 he became professor extraordinarius. His laboratory expenses continued to plague him for more than a decade, but at least he had arrived at a permanent position.27 3. FINDING A NICHE, MAKING A LIVING, C.1800–C.1820 3.1. Hans Christian as a lecturer : Pleasing audiences Having finished his doctorate in 1799, Hans Christian spent most of the next few years looking after Ludvig Manthey’s apothecary while looking for lecturing opportunities. In 1800, he planned to hold a free collegium privatissim for pharmacists every Sunday: Using new discoveries I would eliminate very common prejudices or decide which preparations were the most suitable. For the latter I’d have a couple of the audience do the preparation in our laboratory, and the result of the experiments would give me the opportunity to inform.28

It would be desirable to examine Ørsted’s lecturing practice in much greater detail than I have done here—I have restricted myself to providing some possible pointers. Jan Golinski has argued, with British sources, that experimental natural philosophy contributed to a remodeling of public life as a whole, from the 17th century on29, and we have good characterizations of a number of public lecturers in Britain who may serve to inform our understanding of Ørsted as a lecturer. Joseph “Priestley’s demonstrations and texts were aimed at diffusing factual knowledge among as wide an audience as possible by allowing them to witness, or if possible to replicate, experimental findings. For Priestley, the purpose of this was to provide the population with direct experience of the providential powers of nature in order to liberate them from the ignorance on which corrupt authority was founded.”30 Ørsted certainly also aimed to liberate the populace in the sense of elevating them to a higher cultural consciousness through direct experience of providential powers. But he did not juxtapose it with any corrupt authority. Quite the opposite, Ørsted aligned himself with political power in Copenhagen. His lecturing practice was initially more akin to that of Humphry Davy:

27 28

29

30

Dansk Biografisk Leksikon (n. 4). Mathilde Ørsted (ed.), Breve fra og til Hans Christian Ørsted – Første Samling, Kiøbenhavn: Th. Linds Forlag, 1870, pp. 38–39, 8–9. A few months later he reported doing galvanic-electrical experiments every Sunday afternoon, perhaps in these private colloquia, ibid. pp. 12–13. Jan Golinski, Science as Public Culture—Chemistry and Enlightenment in Britain, 1760–1820, (Cambridge and New York: Cambridge University Press, 1992), 5. Ibid. p. 8.

30

A. HESSENBRUCH The form of public science that Davy constructed was in marked contrast to that favored by Priestley. Far from being invited to share in the production of scientific knowledge by replicating experiments, Davy’s public audience was expected to remain entirely passive, awed by the power of the philosopher and his instruments.31

During his 1801–1803 tour, traveling with his own battery (new and exciting at the time) and thus gaining access to polite society with performances of galvanic experiments, Ørsted learned an important lesson on the role of audiences in experimentation. His experiments at a Madame Henriette Herz’s salon in Berlin32 were to have involved dissected frogs but he could not obtain any and so he did some other small experiments and explained the theory. I find it most awkward to experiment on such occasions where everyone wants to join in with their fingers or at least with their voices, and I am sure that no experiment would have succeeded had I chosen new ones I’d never done before.

Ørsted learned that rehearsals of and procurement for experiments should be done in advance of a public demonstration for maximum effect, and that participation by the audience was disruptive. Ørsted was living through a period in which the salon culture of experimentation was replaced by public lecturing, a topic that deserves much greater attention than I have given it. Some themes pervaded both cultures though, as a brief examination of the London scene of public science performances will show. The notion of preparation in private securing success of a performance in public has been well established for Michael Faraday, Davy’s successor at the Royal Institution in London.33 But there is a further relevant aspect of Faraday’s experimental practice. Faraday construed his public experiments to focus his polite audience’s attention toward the phenomena that he meant to be pregnant with moral and theological meaning. By contrast, William Sturgeon, lecturing to a commercial audience at the London Institution, focused the attention more toward the instrumentation, also because he had designed it and because it was for sale.34 In this respect, Ørsted is very much like Faraday. He had a polite audience, at least in the first two decades of the 19th century, and he focused on the message of the sublime in nature.35

31 32

33

34

35

Ibid. p. 9. Dan Ch. Christensen’s forthcoming Ørsted-biography will analyze the circles in which he moved on the basis of unpublished sources and give us a much better picture than I am able to provide here. It might be well to remember that experimental practice in private and public had not yet hardened into the form that is now so familiar to us. As Iwan Morus has put it, referring especially to the early 19th century: “Experiment, after all, is an attribution that endows a particular event or activity with significance and status. It is not a transcendental category. That a particular way of going on counts as doing an experiment is the outcome of a number of factors ranging from the social status of the performer, through the space where the performance took place, to the outcome of the activity.” Iwan Morus, Frankenstein’s Children—Electricity, Exhibition, and Experiment in Early-NineteenthCentury London, (Princeton, NJ: Princeton University Press, 1998), pp. 10–11. In Copenhagen, the most prominent person to define the new categories of public lecturing was Hans Christian Ørsted, who, as we shall see, indeed contributed to the remodeling of public life as a whole. Iwan Morus, “Different experimental lives: Michael Faraday and William Sturgeon,” History of Science, 30 (1992), pp. 1–28. To my knowledge, he never uses the word sublime but always refers to “the Good, the True, and the Beautiful.” But the sublime covers the meaning entirely satisfactorily.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

31

On his tour of Germany (and France) in 1801–1803 he was very attentive to performance skills: Today I heard Fichte’s lectures for the first time. He has a rare talent for explicating his ideas that renders his lectures doubly interesting. My time was well spent even though I learnt nothing more than some of the tricks by which he attracts and keeps the attention of his audience, prompting the associations in their minds that are difficult to express in words.36

When in Paris, Hans Christian paid attention to lecturing techniques at the École Polytechnique, noticing good and bad practices. What he found particularly worth emulating back home was the students’ hands-on laboratory experience.37 His main encounter on this European tour was with Ritter and the two of them discussed experiments passionately.38 There was much the two could agree upon, first of all the importance of empiricism and the encryption of the sublime in nature. In subsequent years, Ritter was also to provide an instructive example of failure. His use of his own body in experimentation and especially his predilection for achieving trance-like states of mind, eventually undermined his authority as an experimenter. In later years, Ørsted distanced himself carefully from Ritter’s lack of control: He had had a wonderful mind but lacked the requisite self-discipline to steer him clear of illusions.39 Back home in 1804, he secured funds to pay for a collection of scientific instruments and sent out public invitations to lectures on physics and chemistry: “Physics may serve the philosopher as a weapon against the uninitiated who ridicule the holiness of science into which they cannot reach. He may now show them experientially that which heretofore they denounced as the phantoms of unrestricted imagination, putting before their very eyes that which their spirit could not perceive, so that only the most callous can pretend to doubt.”40 The attendance numbers were high, so much so that Hans Christian had difficulties accommodating everyone. C. Hauch described one set of reactions to his 1805–1806 lectures in the following words: He generally started his lectures on a quiet note with simple considerations and observations, occasionally also with explications and determinations of selected

36 37

38

39

40

Mathilde Ørsted (ed.), Breve (n. 28), pp. 38–9, 43. Kirstine Meyer, “H. C. Ørsteds Arbejdsliv i det danske Samfund,” in H. C. Ørsteds naturvidenskabelige Skrifter, edited by Kirstine Meyer, København: Andr. Fred. Høst, 1920, vol. 3, XI-CLXVI, at XVII–XIX. Much attention is being paid to their joint theorizing and friendship with Novalis (Hauch, n. 2, at 122) and less to the fact that unlike Kant, Schelling, or Fichte, they worked with their hands. Hans Christian’s own apologetic explanation of 1826 is that “Ritter had let himself be led astray by his prejudiced imagination. … Winterl was a man of great thoughts but without a keen apprehension of the specific. His experiments, if one may call them so, are without value,” “Ørsted (Hans Christian)”, in Conversations-Lexicon (n. 2), at 523. “Il est vrai que ce n’est pas sans raison qu’on lui a reproché de s’être trop abandonné à des conceptions trop hasardées, et même peut-être extravagantes.” Kirstine Meyer, vol. 2, p. 175. The danger of a loss of authority when one’s body was involved in experimentation was brought to a head in the debates around mesmerism, cf. Alison Winter, Mesmerized: Powers of Mind in Victorian Britain, (Chicago, IL: The University of Chicago Press, 1998). “Indbydelse til physiske og chemiske Forelæsninger,” Reprinted in Meyer, “Arbejdsliv” (n. 37), pp. 78–79.

32

A. HESSENBRUCH expressions, by which he wanted to ensure that he would be understood in the following. Sometimes he lingered on the translation of foreign chemical and physical terminology into Danish. Then he followed a certain chain of thoughts, each part of which related closely to each other and to the basic expressions. To begin with his lecture was only special through a certain keen good sense, but gradually the objects were gathered into larger groupings that were again connected to a larger whole, and that he showed lucidly to the imagination [Anskuelse (German: Anschauung)]. Thus his lecture expanded steadily, like a stream growing from its incorporation of other streams, and eventually it had such a force that the younger in the audience, the receptive to the new and uncommon, the not yet prejudiced, had difficulties withstanding it.41

The style is also visible in the lecture printed as Conversation on mysticism; the format fits the description of streams joining into a large river.42 This crescendostyle lecturing resembles Hans Christian’s preference for dinner parties beginning with cheerful ditties and ending with Schiller’s Ode to Joy. It can also be seen in several of his written papers. It would seem that Hans Christian deliberately aimed to awe. This is of course a good strategy when one needs to attract audiences, particularly fee-paying ones. Something akin to awe or ecstasy is occasioned by the intimate encounter with God’s plan, and Hans Christian’s use of the word “mystic” in the title does give a good sense of the tone in the lecture. In this lecture, he pays attention neither to the dangers that ecstasy poses to self-control nor to possibility that “mystic” may refer to a secret form of knowledge. Later, as we shall see, he would navigate these issues with great care—a care that in retrospect is conspicuously absent in this early lecture. In 1806 he finally secured a position as extraordinary professor at the University, probably on the strength of his successful lectures.43 It is indicative of his increasing emphasis on the role of experimentation that on a second tour in 1812–1813 he found that “owing to the difference of our [his and Fichte’s] views, especially on nature, no very comprehensive communication can take place between us.” By this time “corroboration” was a part of Hans Christian’s vocabulary, the lack of which entails that “unity itself becomes a barren and empty thought leading to no true insight.”44 Hans Christian received the funds for the second European tour on the strength of his lecturing. The public fund (Fonden ad usos publicos) reported that his lectures “have become ever more frequented by members of all estates (Stænder), civil servants including military, artists, manufacturers, tradespeople, etc.”. Hans Christian noted that most of his time was spent either lecturing or preparing for lectures.45 In 1811 he became a lecturer also at the Landcadetacademi, the school of the army. His success brought with it more public support: in 1815, the King 41

42

43

44

45

C. Hauch, “Hans Christian Ørsted’s Levnet” (note 2), at 125–126. Hauch points out that many older attendants were less enthusiastic. It was only published in 1851, but the date given there is 1808, Efterladte Skrifter (note 2), vol. 5, 41–105. “Many learned and distinguished men attended,” “Ørsted (Hans Christian),” in Conversations-Lexicon (note 2), at 526. “[T]he publications and the lectures cleared the way to the professorial position in 1806,” ibid. p. 528. Kirstine Meyer, “The scientific life and works of H. C. Ørsted,” in H. C. Ørsteds naturvidenskabelige Skrifter, edited by Kirstine Meyer, 1920, vol. 1, XIII–CLXVI, at XLVI, the quote is reproduced from Hauch’s biography, cf. n. 2. Kirstine Meyer, “Arbejdsliv” (n. 37), at XXVI.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

33

donated the nation’s largest collection of physical-chemical instruments.46 For the next few years, he barely published, “lecturing 5 hours most days in the winter and 2–3 in the summer.” The lectures continued to open doors to new high-society connections: he held lectures in German for foreign diplomats.47 One of Hans Christian’s main concerns in this period was to find the funds required for running a laboratory; he needed space, the collection of instrumentation needed to be insured, and he had running costs for a “chemical assistant, an untrained assistant, materials for making chemical prepations, oil, spirit, glassware, containers, file, raspe, and similar tools.”48 The fund ad usos publicos was his main source of financial support but he also collaborated with individuals having access to military funds.49 His marriage in 1814 marked a degree of financial independence and in 1815 he was knighted (Ridder af Dannebrog).50 He did publish diligently: a number of monographs and articles in various journals,51 but the craft of public lecturing mattered more for the pursuit of his career. Later, in 1815 he became Secretary of the Royal Academy of Sciences, for which he wrote annual overviews of developments in science and of the Academy’s activities.52 In other words, he contributed significantly to the growing public sphere. But the lecturing remained the backbone of his activities. In 1803 he had stated that moral and physical nature were intimately connected and that without this connection, physics was useless53 and as mentioned he made much of this theme in his lectures. His interest in the sublime must be seen in the context of his need to make a mark with polite audiences, given the cultural dominance of the clergy and the consequent nature of public discourse. It will have been a boon that the intellectuality and high status of German philosophical systems bolstered the argument, but philosophy alone cannot have been the driving force. No doubt, a part of the success of the lectures was his ability to use the experiments to connect with notions of the sublime, and in 1808 he published “On the reason for the pleasure produced by tones,”54 arguing that human recognition of order in nature amounted to an experience of the sublime and divine. This theology of nature allowed him to present himself as a teacher in the mold of the Lutheran pastors: he was showing the presence of God in every little mundane part of the physical world and there was a morality in this fact too. One would have expected nothing else of an academic graduating from Copenhagen University in the 1790s, 46 47

48 49 50 51

52

53 54

Ibid. at XXXII. “Ørsted (Hans Christian),” in Conversations-Lexicon (n. 2), at 532. He held such lectures also in French in 1823–1824, ibid. p. 539. Kirstine Meyer, “Arbejdsliv” (n. 37), at XXXIV. For example Esmarch, ibid. at XXXIV. Hauch, p. 129 (n. 2) Thoughts on the History of Chemistry (1807), the first part of a textbook entitled Science of the general laws of Nature (1809), used both at the University and at the Army Academy, and Recherches sur l’identité des forces chimiques et électriques (1813, original German version 1812); cf. Kirstine Meyer, “Arbejdsliv” (note 37), at XXXI. Ibid. at LIX. The published sources are the basis for my discussion of Hans Christian’s lecturing practice. A more satisfying account could be assembled using archival and manuscript sources. Dansk Biografisk Leksikon (n. 4), p. 198. “Om Grunden til den Fornøjelse, Tonerne frembringe. En Samtale.”, reprinted in Efterladte Skrifter, vol. 3 (n. 2), pp. 67–99.

34

A. HESSENBRUCH

with a public sphere in its infancy, closely watched by the absolutist state, and with the majority of educated people being members of the clergy. Hans Christian’s lecturing not only aimed to convince the University of an expanded role for the natural sciences, but also to curry favor with the political elite. He was occasionally invited to perform experiments at the mansion of Count Schimmelmann, where his knowledge of German and cultivated manners will have stood him in good stead.55 It is interesting to speculate what the noble grandee saw in the young natural philosopher. His fluent German and mastery of German philosophy will have helped, as will his civil demeanor. Hans Christian’s childhood experience with the German wig maker must also have equipped him to at least not reject German culture’s superiority. His lectures will also have been entertaining, and even though the sublime is often recognized as a bourgeois aesthetic in this period, the discourse on the manifestation of the divine in his experimental lectures at least framed the phenomena in a larger and intellectually stimulating way—just the way Faraday had. But there is a further important possible connection. It has been argued that the political elite was combating the guilds and their monopoly of craft knowledge, and that natural philosophy served their purpose as an explicit, articulable, and hence public form of knowledge. Hans Christian’s lectures presented all experimental phenomena as not relying on secret craft knowledge but rather as comprehensible from a theoretical perspective, the perspective of God’s laws. And furthermore, he published textbooks, the very symbol of public knowledge in contrast with the private and secret form of knowledge passed on within the guilds’ master–apprentice relationships.

4. THE 1814 DISPUTE WITH GRUNDTVIG The argument that God manifested Himself in nature was a trope in natural theology in Britain at the same time. The textbook of Paley used in early 19thcentury university education contained the argument by design: that because nature disclosed a pleasing order, it must have been divinely created. John Hedley Brooke has argued that the idea of divine wisdom discernable in nature was attractive equally to Christian apologists and deists. For Christian apologists, it seemed to offer independent proof of a God who had also revealed Himself in the person of Jesus Christ. For deists, the study of nature made the study of the bible redundant. Tom Paine in his Age of Reason, for example, took that argument to its logical conclusion, as summarized by Brooke: Whereas the Bible had been written by men, nature was the handiwork of God. Whereas the Bible had suffered corruption through copying and translation, nature had an indestructible perfection. Whereas the Bible portrays a passionate God, changeable and vindictive, nature shows him to be immutable and benevolent. The biblical revelation had come late in time and been vouchsafed to one nation only. The revelation in nature had been always, and universally, available. In the Bible God communicated with men through magic, in nature it was through their ordinary senses. Paine’s

55

Hauch, p. 127–128 (n. 2).

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

35

triumphant conclusion was that theology was merely the study of human opinions concerning God, whereas science was the study of the divine laws governing nature. Paine, no less than Paley, was confident that nature disclosed a benevolent Creator. Where he differed was in his confidence that people had been equipped with all the natural faculties for working out their own salvation.56

To Hans Christian, his way of pleasing polite audiences was in tune with Lutheranism and as such with the general role of the Danish teacher-pastor, but his discourse also rang the alarm bells with others, just as deism did in Britain. The populist preacher, N. F. S. Grundtvig, challenged precisely this (at the time of the Restoration, a Europe-wide wave of antiliberal retrenchment57): it is said that I talk and gesticulate following blindly the impulses of the moment and quixotic whims without recourse to calm reason. The impulse of the moment, when in accordance with God’s written word, is worth more to me than all the world’s clever reasoning. To take apart a clear truth bit by bit, to pick the meat off and then to link the bones up with wire seems to me an indignity. It is disgusting to see how Reason boasts of such a heroic deed, as if it had generated the truth by thus anatomizing it, as if it were the Lord itself.58

Grundtvig had no truck with reasoning, insisting instead that truths are clear and not in need of explication. Clearly, the genre of philological analyses of the Bible that had forced the Church into rearguard action, was a main target but he railed against all forms of Reason. He argued also that the “physical and mathematical sciences were much more likely to entertain those who wished to forget God than the historical ones.”59 Hans Christian bristled; his book, Against the Large Accuser of 1814 is a volume full of arguments against “the obscurantist”60 Grundtvig’s “prejudices, called into being by his wild desire to find evil everywhere and by his hatred of light.” Hans Christian meticulously went through a series of Grundtvig’s publications, showing what he took to be a lack of comprehension. He is particularly scathing about Grundtvig’s reading of Schelling: the person, with such a lack of philosophical education [Dannelse, German; Bildung] as that exhibited throughout by Grundtvig, who tries to read Schelling, must get into a similar frame of mind as the mechanical artist who attempts to deduce the foundations of his craft from reading Laplace’s Mechanique celèste. […]61

Obviously, Hans Christian is pulling rank: he’s a university professor explaining that the mere pastor does not understand. He also terms the derision of philosophers a “triumph of the common mob.”62 Haughtiness was not foreign to Hans

56

57

58 59 60 61 62

John Hedley Brooke, Science and Religion—Some Historical Perspectives (Cambridge: Cambridge University Press, 1991), pp. 193–194. The word deism is also used by Ørsted’s first biographer, Hauch, p. 173 (n. 2). The Ørsted brothers read Payne with enthusiasm as boys, cf. Anders Sandøe Ørsted, Af mit Livs (n. 22), p. 18. Robert Fox, “The rise and fall of Laplacian physics,” Historical Studies in the Physical Sciences, 4 (1974), pp. 89–136. Quoted in H. C. Ørsted, Imod den store Anklager, Kiøbenhavn: Andreas Seidelin, (1814), p. 50. Quoted in ibid. p. 88. Ibid. p. 50. Ibid. p. 63. Ibid. p. 52.

36

A. HESSENBRUCH

Christian. But there is a sense in which the mere pastor is the authoritarian, not the university professor. After all, Grundtvig talks about the truth not as something to be worked out but as a given and not negotiable. Truth just is, and remains clear to him even in the face of disagreement with others. If Grundtvig asserts that something is the truth, and that he can see a clear basis for it in the Bible, then there is nothing left to discuss. By contrast, Hans Christian does not posit an authority in one text or in one person, such as himself qua university professor. Rather, he posits it in a kind of method, Reason, accessible to all human beings with the one proviso that they be properly educated/cultivated (the debate took place in 1814, the year that primary education became compulsory): it is insufficient to have a humanly correct idea of the Good, the True, and the Beautiful in order to apply it in life. One has to always compare the objects with the ideas. Mostly this is easy and may be done using tact (this latter may be natural and it may be developed through practice). Often, however, the comparison is more difficult, in which case one has to either deploy the most keen reflection or confide in men whose understanding and probity may be trusted. As surely as the spirit ought to diligently turn inwardly its undivided attention to strive through devout exaltation for those blissful moments of rapture, as surely as these moments must warm and illuminate all other parts of life; just as sure it is that in order to see clearly in the finite circle within which we have been placed by God, we dare not reject the analytical art of investigation.63

I would like to make three points about this passage. For one thing, Hans Christian is careful to show off his religious credentials, implying that he is not in any confrontation with the Danish state church. Secondly, the more education one receives, the more one will be enabled to discover truths oneself. With the broad education of the populace, more and more individuals will be increasingly enabled to judge for themselves. Whereas Grundtvig’s attitude to authority is that it is lodged unequivocally and unanalyzably in a text, the Bible, Hans Christian claims that it can be lodged neither in a particular text64 nor in an individual, with the exception of those who have attained a certain level of cultivation and probity. And as such the authority lies not in the individual but in the attainment. Thirdly and finally, it is interesting that Hans Christian uses the word “tact,” not discipline and not skill. He was a man of neither the military nor industry. The issue is neither submitting oneself to an authority, nor deskilling. He was a man of the bourgeoisie, the typical training of which was music. Learning music is to become cultivated, to possess self-control, and to appreciate the sublime. Tact refers to control of one’s passions, pointedly opposed to Grundtvig’s impulsive and boisterous behavior, and to individuals out of free will choosing to behave appropriately. He never refers to any rigid social hierarchy such as Estates or station, only to acquirable behavior such as tact. And tact is concerned with more than manual dexterity, unlike skill, namely a socially appropriate form of cultivation.

63 64

Ibid. pp. 53–54. A few pages earlier he had explained with the example of a judge, presumably inspired by a conversation with his brother, who may well have an excellent law book to aid them, but if he has no knowledge of the actual case at hand, he remains helpless. The judge needs three things: knowledge of the written law, understanding of the case at hand, and good judgment, ibid. pp. 48–49. Authority cannot be lodged in the written law alone.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

37

Hans Christian talked the talk but also walked the walk by reprinting in extenso those of Grundtvig’s texts that he engaged with, peppering his arguments with references, inviting the reader to go and judge for himself.65 And then he uses Grundtvig’s own source of authority, the Bible, to underscore the legitimacy of this method: in the words of Paul: “Try all and keep the good.” Hans Christian must have been confident that the reader be able to deduce that he was right and Grundtvig wrong. Hence, I presume, the prominence of the 8th commandment of Luther’s catechism on the front page: You must not tell lies about your neighbor. That is: We must fear and love God, so that we will not deceive by lying, betraying, slandering or ruining our neighbor’s reputation, but will defend him, say good things about him, and see the best side of everything he does.

The encounter with Grundtvig thus marks a change in Hans Christian’s language. In the 1808 lecture, he had still talked about the secrets of nature and about mysticism with some indication of fascination. After 1814, he emphasized the transparency of scientific truths. Awe and ecstasy have been carefully expunged. Rather, Hans Christian’s language is better described by the Kantian sublime: Reason is at first confounded by the experience of the very small, large, fast, or seemingly unclassifiable. But crucially, Reason recoils and copes. The encounter with the sublime only initially makes one feel small. Autonomous reason is in fact civilized and educated through the challenging encounter with the sublime.

5. POWER AND LEVERAGE 1820–1840; AND DECLINE 1840–1851 5.1. Hans Christian’s institutional power and influence In 1815 he became secretary of the Royal Academy of Sciences, in 1817 professor ordinarius and faculty board member. The 1820 discovery of electromagnetism brought fame and increased stature. He finally managed to garner enough financial support to put his laboratory on a secure footing. He broke into the innermost circles of power: the Crown Prince attended his lectures, and later, as King Christian VIII, allowed Hans Christian liberal access to him, for instance for advice on academic appointments.66 In 1825–1826 he became chancellor of the University (with two further stints in 1840–1841 and 185167), in 1819 also chair of a government commission mapping minerals on the island of Bornholm. With his increased stature he initiated and shaped two educational novelties, the Society for the Advancement of Science (founded in 1824–1825),68 and the Polytechnic Institute (1829). The University had resisted governmental pressure for reform in the 65 66 67 68

Ibid. pp. 55, 83, 98, 133, 137. This list of pages is not exhaustive. Hauch (n. 2), p. 145. When a faculty of science was finally accepted. Meyer argues that Hans Christian was inspired by the funding through subscription of the London Institution that he visited in 1822, but the Danish version sent teachers to teach throughout the country and did not have a central building with a lecture hall in the capital. Kirstine Meyer,

38

A. HESSENBRUCH

late 18th century and still did not welcome the natural sciences. Its governing body gradually accepted Hans Christian personally, but not until 1850 did it accept a faculty of sciences.69 Much of his efforts in the last decades of his life were spent as an institutional politician. In 1828 he picked up a suggestion for a new school for craftsmen modeled on the German Gewerbeschulen. He deftly played University politics for 18 intense months to have the Polytechnic Institute founded, modeled on the Parisian École Polytechnique.70 The economic climate was still not conducive to institutional innovation: the Danish government budget was reeling from the after-effects of the Napoleonic Wars and the 1813 national bankruptcy.71 However, the increasing specialization of the civil service generated a demand for technical education.72 The 1820s saw an increase in journals communicating technical knowledge, revealing the general growth of interest in such matters.73 A similar development took place within the military and an institution for military-technical education, especially artillery, was founded in 1828—essentially at the same time. Still, the negotiations were difficult, and only by the pooling of resources, including the offer to use parts of Hans Christian’s own home for teaching purposes, by diligence, and by concessions to the many diverse interests involved, did he eventually succeed.74 Michael Wagner has argued that a part of the development was driven by a desire to break the grip of the guilds’ monopoly on technical knowledge. The enlightenment ideal of making technical knowledge public and accessible to all was to undermine the grip of the guilds, within which knowledge was passed from master to apprentice only and access to it was impossible for anyone else. During the life span of Hans Christian, the guilds were in fact effectively undermined.75 Hans Christian’s perspective of the general theories from which all specific knowledge could be derived, was exactly the kind of bird’s eye view that undermined the guilds’ monopoly: The essence of the theoretical concept of knowledge was that technology was exclusively regarded as a form of knowledge expressed in language—a language that partly had to be created to encompass the useful knowledge of nature, technology, and society.

69

70

71 72 73 74 75

“Arbejdsliv” (n. 37), at XCVI. Also, according to Hauch (n. 2), p. 178, Hans Christian was a member of the Royal Institution, not the London, which would make more sense, given the comparison with Faraday above. The Society also co-organized an Industrial Exhibition in Copenhagen in 1836 highlighting education of håndværkerey, Sunday schools, and travel grants for håndværkere and manufacturers, Ida Haugsted, “Industriudstillingen 1836—Design, teknik, håndværk,” in Krydsfelt (n. 2), at 171. Michael F. Wagner, Det polytekniske gennembrud (n. 19), p. 122; Ole B. Thomsen, Embedsstudiernes universitet- en undersøgelse af Københavns universitets fundats af 1788 som grundlag for vores nuværende studiestruktur, 2 vols., København: Akademisk Forlag, 1975. Michael F. Wagner, Det polytekniske gennembrud (n. 19), pp. 185–303. Bruno Belhoste et al. La Formation Polytechnicienne : 1794–1994, Paris: Dunod, 1994 Michael F. Wagner, Det polytekniske gennembrud (n. 19), p. 291. Ibid. pp. 124–125. Ibid. p. 182 Ibid. p. 294. Ibid. pp. 113, 121.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

39

The point was to transform the knowledge that traditionally had been the exclusive terrain of the artisanry into a systematic generally educating/cultivating form of knowledge. This technical knowledge was developed in the shape of a new academic discipline, polytechnics, sustained by and transmitted through societies and institutions with no direct contact to labor and production, but with access to the public sphere.76

According to Wagner the people involved in the making of this new discipline were primarily civil servants, academics, a few noblemen, and assorted craftsmen and industrialists. The civil servants in particular and noblemen such as Schimmelmann who sat at the very top of the administration were exactly the kind of people who wanted a bird’s-eye view. It is also precisely the viewpoint that was to make Hans Christian look reactionary in 1850. Michael Wagner has further argued that Hans Christian’s primary concern was to create an institutional home for the natural sciences. He had failed for a long time to have the University accept a faculty of sciences, and the Polytechnic Institute constituted the best conceivable compromise.77 The University’s governing body played an active role early on, by taking charge of the committee for creating the new institute, but once Hans Christian had been placed as chairman of the committee, they left the rest to him.78 He also shaped the curriculum of the new Polytechnic Institute, placing great emphasis on conveying a general understanding of scientific laws. The artisanal aspect of technical knowledge was less prominent. Considering his own predilections, but also the need to stay on the right side of the University’s governing body, this is hardly surprising. The tremendous status of the École Polytechnique throughout Europe, with its highly theoretical approach, will also have influenced him. The Polytechnic was placed under the Ministry of Culture, and it is indicative that this ministry was also in charge of Board Schools, municipal primary schools, grammar schools, and the University.79 One might imagine for instance that an engineering school would have been placed under the Ministry of the Interior in charge of manufacturing and industrial matters, but that was not the case. In other words, the purpose of the Polytechnic was to be coextensive with the purpose of Hans Christian’s life in general: it was another brick in the edifice of general education/cultivation of the populace. As Michael Wagner has put it: The romantic concept of polytechnics turned out to be a suitable platform for the shaping of a political compromise in the educational sphere.80

In his new role as a political animal within the educational corridors of power, Hans Christian began to publish on new topics.

76 77

78 79 80

Ibid. pp. 109–110 Ibid. p. 220; also p. 124 and Kirstine Meyer, vol. 3 (n. 37); and Jørgen Broberg Nielsen and Eivind Slottved, “Fakultetets almindelige historie,” in Københavns Universitets historie 1479–1979—Bind XII—Det matematisk-naturvidenskabelige Fakultet, København, 1983. Michael F. Wagner, Det polytekniske gennembrud (n. 19), p. 293. Ibid. p. 382. Ibid. p. 294. Wagner accepts and uses the dismissive vocabulary of the subsequent generation: romantic had the overtones of the lack of self-discipline that was pinned on Ritter. Ørsted would have flinched at this attribution.

40

A. HESSENBRUCH

6. EXCURSUS: ANDERS SANDØE ON FREEDOM AND MORALITY Having followed Hans Christian’s career I have ignored his brother Anders Sandøe. But once Hans Christian’s interests turned to the general role of science in society, Anders Sandøe provided an important resource for him, so we will now turn to his work in legal practice and jurisprudence. Not surprisingly, Anders Sandøe’s development took a different turn, although both lost their youthful enthusiasm for philosophical systems once they encountered the practicalities of, respectively, being a lecturer and a judge. But whereas Hans Christian retained a fondness for the transcendental in his conception of the Good, the True, and the Beautiful, Anders Sandøe became thoroughly pragmatic. Anders Sandøe rose even higher than Hans Christian. In 1810 he had become a High Court judge, which he left in 1813 to join the central administration. In 1825 he became generalprokurør, the government official in charge of drafting new legislation. His thinking here resembled that which he had worked out as a judge: grand comprehensive systems ought not to be foisted upon a complex reality. The organically grown economic and social spheres were much better served by a correspondingly unsystematic law, reflecting the mores of the time; in other words by positive law.81 Contemporary German legal scholars such as Savigny,82 engaging with developments in France, debated codification and Anders Sandøe read them extensively, reporting on his readings in print. He turned decidedly against a natural law with norms overruling existing positive law, and emphasized instead precedence, just like English law: in legal practice one ought to interpret existing law to intuit how to expand it to cover more of social life. Danish penal law, written in 1683, was out of touch with 19th-century society. For instance, it knew of no such crimes as fraud and embezzlement, and because based on lex talion it had no category of light punishment (it was based on the Ten Commandments83). In 1800 this led to a great many pardons because contemporaries found the heavy-handed punishments excessive. In his gold medal essay Anders Sandøe had made a transition to Fichte’s system, despite its “invasive character and its requirement for a heavy-handed police.” Around 1800 he became convinced that Fichte’s system, far from shoring up freedom through a rigorous social system, actually destroyed it. It would amount to terror if all citizens enjoyed perfect security if that security was built on fear and coercion, while “justice, love, trust, and honor were forced into exile.”84 He began to reject all philosophical systems.85 Anders Sandøe worked hard to make punishment fit the crime, and because of the many new forms of crime, eventually much new 81

82 83

84 85

For example Ditlev Tamm, Fra “lovkyndighed” til “retsvidenskab”—Studier over betydningen af fremmed ret for Anders Sandøe Ørsteds privatretlige forfatterskab, København: Juristforbundets Forlag, 1976, 17. A similar development from natural law to positivist law took place elsewhere in Europe, ibid. p. 23. Ibid. p. 37. Troels G. Jørgensen, De Ørstedske Straffelove, Copenhagen: Gyldendalske Boghandel Nordisk Forlag, (1948), p. 10. Anders Sandøe Ørsted, Af mit Livs (n. 22), p. 117. Ibid. pp. 78–79.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

41

legislation had to be written. His most productive period was 1833–1841 when, as Generalprokurør (government official in charge of drafting new legislation), he wrote new laws on bodily harm that for the first time stated general rules about attempts and assistance; on theft and fraud; on false testimony and perjury; and on arson.86 He contributed a great deal toward the replacement in Denmark of law suited not to a society with arbitrary and violent power but with the sort of micropower that Foucault has described.87 In particular, he was not particularly concerned with the disciplining of bodies in the sense of the meticulous ordering of mundane life introduced in the prison, but rather with the nurturing of morality that Foucault referred to as inculcation. Indeed, the mature Anders Sandøe was the kind of progressive icon that caused Foucault’s ire. Looking back in 1826, Anders Sandøe explained the legal thinking that he had by then rejected: Around 1800, systems of law, especially criminal law, were characterized by a serious, strict, even unyielding and hard philosophy. Their aim was strict abidance by the law, ordered through clear and distinct concepts; an aim that yielded to no other consideration. Penal law merely drew exact boundaries and discussed what was counter to the spirit of the law, the purpose of which was to eliminate any arbitrary actions by judges. Defendants’ mental state was considered only in the context of the question whether the crime had been deliberate or not. In this respect at least, the two most influential philosophers of the time, Kant and Fichte, agreed. Both assumed a keenly distinguishing and limiting law meting out with mathematical precision a law that absolutely had to be followed regardless of all other considerations, even ethical ones. Their motto was fiat justitia et pereat mundus. Any violation of this law had to be punished severely and according to the letter of the law, regardless of the degree of ethical viciousness.[…]The two doyens of criminal jurisprudence, Feuerbach and Grollmann, agreed, and also emphasized the importance of a legal system curtailing arbitrary judgments in court and independent of ethical concerns.88

As a judge, Anders Sandøe may well have resented the merely mechanical role that such legal philosophies assigned to the judge. Leaving a measure of decision-making to judges who understand the contexts of each particular case, he argued, was preferable to a centrally defined law, shoehorned into every single case. Whereas Kant and others had thought primarily in terms of a central power safeguarding the rights of citizens against the abuses of law lords or other local potentates, Ørsted emphasized the superior results by devolved power, where decisions could be made locally. Anders Sandøe used the phrase “nature of the case at hand” (Sagens Natur) extensively, which has become celebrated by historians of Danish law as his particular approach to law.89 The main difference between Kant and Ørsted lies in the ethical behavior of local power. Kant assumed that local power would inevitably be abused, and only the curtailment of discretion could limit that abuse. Anders Sandøe assumed that local judges could behave ethically. In an 1807 essay on The Relationship between Government and Religion, he emphasized that legal order cannot be built on a foundation that has no 86

87 88 89

Ditlev Tamm & Jens Ulf Jørgensen, Dansk retshistorie i hovedpunkter—Fra landskabslovene til Ørsted – II – Oversigt over retsudviklingen, København: Universitetsforlaget i København, (1973), p. 51. Michel Foucault, Surveiller et punir—naissance de la prison, Paris : Gallimard, (1975). Quoted in “Ørsted (Anders Sandøe), in Conversations-Lexicon (n. 2), at 555–556. Ditlev Tamm and Jens Ulf Jørgensen, Dansk retshistorie i hovedpunkter (n. 86), p. 173.

42

A. HESSENBRUCH

recourse to morality. The point was that the legal system can only function when individuals within it, be they judges or witnesses, behave ethically. Hence a major task of building a just society must be to inculcate morality into every single citizen (Foucault’s theme).90 Anders Sandøe not only theorized in private about such matters but also worked diligently to diffuse his ideas. It has been said of him that as a descendant of a long line of pastors, his need to enlighten and guide was inherent,91 but the point is rather that in a society with delegated responsibilities, one needs an active public sphere. The kind of society that Anders Sandøe envisioned required that he publish and teach prodigiously—and of course such activities simultaneously enhanced his stature. In 1809 he joined the board of directors of the pastors’ seminary and lectured in canon law, supervised law students at the University, and he was extremely productive, the tone of his writing being explicative, even didactic. Many of his early writings were published in journals that he had himself started and/or of which he was the editor, e.g. Juridisk Maanedstidende, Juridisk Arkiv, Nyt juridisk Arkiv, and Juridisk Tidsskrift (Juridical Monthly Newspaper, Juridical Archive, New Juridical Archive, and Juridical Magazine). In addition to a long stream of comments on court decisions and smaller theses, he published three volumes of comments on Danish and Norwegian private law (1804–1812), a volume on the concept of theft in penal law (1809), and a series of foundational volumes on penal law and the role of proof, under the umbrella name of Eunomia (1815–1822). He is the most prolific contributor to the budding public discourse on law in Denmark. Publishing did not hurt his career: in 1810 he became Superior Court judge, a position he left in 1813 for one in the central administration.92 He simultaneously advanced his own career and worked towards the realization of the good society: the rule of law shaped to meet the needs of the populace93 and the cultivation of the populace required for the functioning of the legal system: The state ought to contribute its utmost to organize the most appealing, noblest, and happiest external conditions that harmonize as much as possible with the human being’s physical and moral interests. The governmental plan ought to include education to true morality and religion, science and arts, the domination of the rough natural forces, and the most purposeful application of the mild natural forces.94

The grip of the law expanded, both with an increasingly detailed jurisprudence and also with a growing civil service capable of enforcing the law’s expansion into 90

91

92 93

94

“Ørsted (Anders Sandøe), in Conversations-Lexicon (n. 17), at 560; also Anders Sandøe Ørsted, Af mit Livs (n. 22), pp. 118–120 Troels Georg Jørgensen, Anders Sandøe Ørsted—juristen og politikeren (København: A. FrostHansen, 1957), p. 287. Dansk Biografisk Leksikon (n. 4), pp. 189–190. “I have little respect for so-called legal philosophies consisting of dead meaningless form and without any connection to real justice which after all is supposed to be the scopus jurisprudentiæ. The day-to-day law that serves to advance and maintain legal order is just in my opinion, and I know no other source for determining rights.” Juristisk Arkiv, no. 4 (1805), p. 108; quoted in Troels G. Jørgensen, Anders Sandøe Ørsted som Dommer (København, H. Hagerups Forlag, 1928), p. 20. Quoted in Carl Henrik Koch, “Ørsted og Striden om Viljens Frihed”, in Anders Sandøe Ørsted 1778–1978—Foredrag i anledning af 200-året for Anders Sandøe Ørsteds fødsel, edited by Ditlev Tamm (København: Juristforbundets Forlag, 1980), 87–121, at 90.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

43

the nooks and crannies of society. For example, the rights and duties of landowner and copyholder were specified.95 The state cannot impose religion and morals. But it can arrange for institutions aiming to nurture the citizens’ religious and moral traits (this is the project that Foucault found most invasive and he preferred the term “inculcate” to “nurture”). Law and order can only be maintained if this project succeeds. For example, deterrence may well work in the case of planned crime, but crime committed in anger is unlikely to be deterred by the imagination of punishment. A moral disposition is more likely to restrain sensual cravings and criminal inclinations. An oath taken as a witness can fulfill its legal purpose only if the probity of the witness is beyond doubt. Anders Sandøe explicated his difference of opinion with Kant, who had talked about the moral freedom of the rational human being, that is the ability to act on his/her own accord (unlike a machine). Indeed, this freedom has to be taken for granted for a legal system to function (culpability gained general acceptance in law in this period96), but the legal system has to support freedom in a much more expanded sense. Civil freedoms have nothing to do with the moral freedom that according to Kant distinguishes the rational human being from the machine.97 Anders Sandøe was loath to yield where essential concepts such as culpability and control of one’s passions were concerned. Medical expertise claimed a right to determine criminal responsibility: In 1824 the medical doctor F. G. Howitz argued that an insanity defense, to use a modern term, could well be reasonable. Anders Sandøe rejected the notion that insanity could be a fact, and that the decision should be left to the discretion of the judge98:

95

96 97

98

Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 49. The timing of Anders Sandøe’s activities tallies well with the 1813–1815 shifts in Europe with regard to serfdom, clerical courts, and noble courts. In France, for example, new legal codes were introduced: Code Civil 1804, Code Commercial 1808, Code Criminel 1813. cf. Isser Woloch, The new regime: transformations of the French civic order, 1789–1820s (New York: Norton, 1994). Ditlev Tamm, Fra “lovkyndighed” til “retsvidenskab” (n. 81), p. 25. Simon Schaffer, “Enlightened automata,” in Sciences in Enlightened Europe, edited by Jan Golinski, Simon Schaffer, and William Clark (Chicago, IL: University of Chicago Press, 1999), pp. 126–165. Anders Sandøe Ørsted, Af mit Livs (n. 22), 122. Also: Dansk Biografisk Leksikon (n. 4), 191; Carl Henrik Koch, “Ørsted og Striden om Viljens Frihed,” in Anders Sandøe Ørsted 1778–1978—Foredrag i anledning af 200-året for Anders Sandøe Ørsteds fødsel, edited by Ditlev Tamm, København: Juristforbundets Forlag, 1980, pp. 87–121. Koch sides with Howitz, and as Charles Rosenberg argues: in late 19th-century USA and now, forensic liberalism “assume a deterministic stance, conservatives a less deterministic one; Charles Rosenberg, The trial of the Assassin Guiteau— Psychiatry and Law in the Gilded Age, Chicago and London: The University of Chicago Press, 1968, xv. One could indeed map Howitz, Koch, and Anders Sandøe along a left-right political spectrum. (The theme of psychiatry in the context of the insanity defense recurred in US history (and presumably elsewhere); cf. James C. Mohr, Doctors and the Law—Medical Jurisprudence in Nineteenth-Century America (Baltimore and London: Johns Hopkins University Press, 1993). Anders Sandøe championed private property too, which, although it may have seemed liberating in the Ørsteds’ youth, this was by no means obvious by the time of their death. Not surprisingly, Karl Marx was less than impressed with Anders Sandøe’s autobiography: “a colossal state haemorrhoid,” cf. Karl Marx to Friedrich Engels, 22 May 1857, in Marx and Engels Complete Works, vol. 40, Moscow, 1929, p. 132. “an enormous political hemorrhoid” in Letters of Karl Marx, ed. 5. Podover, New Jersey, 1979, p. 117.

44

A. HESSENBRUCH where the supposed mental disruption isn’t somatic, doctors are in general no better than other reasonable and enlightened men at judging. Certainly the judge is better equipped to judge this than doctors proceeding only from the archival records.99

Eventually, Anders Sandøe advocated a compromise by reducing the severity of the penalty in cases of doubt, again diverging from Kantian rigor.100 In 1826 Anders Sandøe also debated with Grundtvig, who, on the basis of the oath taken when sworn in as a pastor, had ferociously attacked a university professor’s rationalist reading of the Bible.101 In an article on “Is Danish canon law in need of thorough change” in Juridisk Tidsskrift (Juridical Magazine), Anders Sandøe discussed the meaning of the swearing-in oath. In it he expressed the opinion that it was quite reasonable to have the most intelligent and most learned of any church community be in charge of its doctrines, and that they adapt it to the times taking into account the most recent science.102 Of course he addressed also the legal context, arguing basically for freedom of speech and against the role of Lutheran pastors, or indeed anyone, as censors.103 Anders Sandøe’s thorough consideration of the role of morality for a good society provided a resource for Hans Christian. If the encounter with the sublime somehow improved morality, then it had a role to play in the building of a new civil society: The utility of science will increase with humans’ trust in it. They have to be permeated by the conviction that beautification and ennobling of human life is founded in true fear of God but also in arts and sciences, by the conviction that they work for their benefit even of the ones who don’t themselves possess them. The degree to which this takes place is the degree to which the Realm of Reason will spread upon earth.104

7. HANS CHRISTIAN PICKS UP THE THEME OF DANNELSE In the early 1830s the infrastructure for schooling of the general population, including the peasants, had been completed and illiteracy was eliminated. By law, the distance to school could not be more than one quarter of a Danish mile = appr. 2 km.105 The subjects were religion, the three Rs, song, and if possible physical 99 100 101

102

103

104 105

Anders Sandøe Ørsted, Af mit Livs (n. 22), 183. Ibid., 186–187. N. F. S. Grundtvig, Vigtige Spørgsmaal til Danmarks Lovkyndige, Kjöbenhavn: Wahl, 1826, esp. pp. 12–13. Grundtvig’s stance was that authority should be lodged with people like him, and as such generally against the academic expertise, the “exegetical popism.” It reacted against ever more rationalist Bible-analysis from academic theologians, and especially against attempts to ground the teachings of the church in rationalist readings. He insisted on the role of common sense, giving as an example also heliocentrism as an example of the academic teachings violating common sense. Sune Auken, “Naturen som tegn—Om Grundtvig og naturvidenskaben,” in Krydsfelt—Ånd og natur i Guldalderen, edited by Mogens Bencard, (København: Gyldendal, 2000), pp. 214–223, at 219 and 221. Leif Grane, “Ørsted og kirkekampen i 1820-rne,” in Anders Sandøe Ørsted 1778–1978—Foredrag i anledning af 200-året for Anders Sandøe Ørsteds fødsel, edited by Ditlev Tamm, (København: Juristforbundets Forlag, 1980), pp. 122–147, at 125. He got into political trouble as a result and was ordered to curtail his public scholarship. “Ørsted, Anders Sandøe”, Dansk Biografisk Leksikon, Sekstende Bind, Woldbye-Aastrup, (København: Gyldendal, 1984), pp. 192–194. “Ørsted (Hans Christian),” in Conversations-Lexicon (note 2), at 541. Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 159.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

45

education. Reading was to include history and geography. The aristocraticcosmopolitan form of education had now been entirely replaced by the academic one. The new watchword was Dannelse (education/cultivation): “academic education had become dominant and the educated were now dominant.”106 Hans Christian did use the term in the 1814 debate with Grundtvig, but not until the 1820s did it become Hans Christian’s single most important word. One important instance is an 1824 essay called “Natural Science considered as a fundamental part of the human being’s Dannelse.”107 The 1824 argument was that it strengthens the love of Reason/God and in so doing develops the “hidden forces” of the spirit. The experimental aspect of science also teaches us how to arrive at secure knowledge. The context for this was not only the sort of challenge that Grundtvig represented but also public challenges to his report on the prospects for mining on the island of Bornholm. (Later, in 1836, having been for seven years Director of the Polytechnic Institute producing some half a dozen candidates every year, the role of Dannelse received a new twist: he now emphasized the prospect of social mobility for educated/cultivated individuals.108) At a University celebration of the King’s birthday in 1826, he argued that “The Enlightenment of the People brings Good Fortune to the Sovereign.”109 There are at least two interesting aspects of it. One is that he felt the need to counter a claim that educating/cultivating the populace endanger the state. By contrast, he argued, it increases the sovereign’s power and makes him safer. This kind of language is one of a radical faced with resistance to social mobility. It is certainly not the language of a threatened social elite. The other interesting aspect is that the article was written shortly before he was involved in the foundation of the Polytechnic Institute. Prior to 1826, Hans Christian had only ever once mentioned utility in the sense of applied science, technology and engineering; in 1809 when he had mentioned it in passing and explicitly stated that its importance was at best secondary.110 A trickle of articles on applied science began to appear from 1817 onwards; on wine-making from fruit, mining on Bornholm, and the chemistry of trees. The trickle became a flood in 1829, with the foundation of the Polytechnic Institute. From then on, the question of utility became almost as important to his vocabulary as Dannelse. Hans Christian’s discourse on the nurturing of morality was well suited to sway the mandarins at Copenhagen University, but it was increasingly out of touch with the extramural world, where both nongovernmental manufacturing and nongovernmental political organizations assumed greater roles. The 1830 revolution 106

107

108

109

110

Ibid., 164. Dannelse has the same connotations as the German Bildung, and may be translated as the process of becoming educated, civilized, and cultured. “Naturvidenskaben betragtet som en af Grundbestanddelene i Menneskets Dannelse,” Efterladte Skrifter, vol. 5 (n. 2), pp. 129–142 (original in Nyt Aftenblad, 10 January 1824). “Ingen Rangstrid mellem de forskjellige Stænder” (Against ranking amongst the different estates), Efterladte Skrifter, vol. 7 (n. 2), pp. 97–102, original in Dansk Folkeblad, 1836. The immediate context of this article was the new assembly of the estates and the heated negotiation of hierarchies there. “Folkets Oplysning Fyrsten Heldbringende,” Efterladte Skrifter, vol. 5 (n. 2), pp. 143–166 (original in Nyt Aftenblad, 18 March 1826). “Den almindelige Naturlæres Aand og Væsen”, reprinted in Efterladte Skrifter, vol. 5 (n. 2), pp. 106–128; at 115–116.

46

A. HESSENBRUCH

in Paris inspired reforms, resented by close allies of the Ørsteds. Count Schimmelmann stated that the new French development was based on the “dangerous principle that the power of the sovereign rested on the will of the people.”111 It was decided to introduce four geographically dispersed popular assemblies to debate government and politics. The landed gentry was to constitute much of the assemblies’ membership.112 The public sphere was growing but the Absolutist government still had not let go of the reins. The assemblies met behind closed doors, and censorship continued. But piecemeal public debate grew, for example in 1835 the government’s budget was published for the first time, as indeed it was in Prussia too. Many newspapers and journals came into being, one of which, Maanedsskriftet for Literatur (Monthly Journal for Literature), Hans Christian had a hand in founding.113 Both brothers engaged actively to counter censorship in the mid-1830s, at the cost of royal displeasure and diminished direct influence upon the government.114 Humor entered the theatre world, contrasting markedly with the solemn polite style that the Ørsted’s had internalized.115 The biting sarcasm of Søren Kierkegaard and his playful duplicity, hiding behind a thicket of noms de plumes, provide a glimpse of a new style of public discourse very different from the one the Ørsteds knew and mastered.116 The University changed its nature from primarily an education for pastors and jurists to provide general academic education/cultivation, now the “requisite for a prominent social position in terms of influence on public opinion, spiritually, or politically.”117 By the 1830s, academic education became the dominant form and the background of those in power. And at the same time that the prominence of University professors increased steadily, the discourse also generally shifted from aesthetic to social issues.118 The growing rural affluence fueled a larger demand required for manufacturing, such as brick works and artificial fertilizers. Other manufacturers took advantage of the cheaper rural labor pool, such as dyeing or weaving manufactures. Increasingly, local iron foundries supplied the market with swing ploughs. But the level of mechanization remained low. Even the capital had only 15 steam engines as late as 1839. The Copenhagen industry employed in total some 6,600 individuals and mostly in a putting out system. Steam engine power was doubled in the 1840s, but the social transformation now associated with industrialization was still to come by the time of Hans Christian’s death in 1851.119

111 112 113 114 115 116

117 118 119

Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 185. Ibid. p. 52. Hauch (n. 2), p. 142. Hauch (n. 2), pp. 143–145. “Serious” is a common term used to describe Hans Christian, e.g. Hauch (n. 2), p. 148. Joakim Garff, SAK—Søren Aabye Kierkegaard—en biografi, (København: Gads Forlag, 2000), esp. pp. 54–91. The public sphere changes match those in Germany in this period. Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 162. Ibid. p. 163. Ibid. p. 320–324. Hans Christian seems to have been less interested in the boundary between man and machine than contemporary British natural philosophers, for whom the machine and especially the steam engine was much more present.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

47

Increasingly, with the proto-industrialization of Danish society, Hans Christian’s audience changed from a polite one interested in displays of the sublime and moralistic arguments to a commercially oriented one interested in skills and utility. Accordingly, he changed his topics towards the application of science to handicraft (Gewerbefleiss), housekeeping, agriculture, machine construction, and extraction of the most useful animal and vegetable ingredients. He neither changed his apodictic tone nor abandoned the notion of the sublime, but aimed instead to convince his various audiences that an understanding, as a later age would put it, of the fundamental laws of nature could be utilized in guiding technological and engineering activity. His scientific publications in this period also make a clear turn towards the new demands. Many publications relate to new instrumentation, for example an electrometer,120 an instrument for measuring capillarity,121 strong magnets,122 and an instrument for measuring depth at sea.123 Soon after he heard of Grove’s element, a much improved battery, he collaborated with the Royal Porcelain Factory to prepare platinum cylinders with the best possible cost–benefit ratio.124 Tensions erupted both within the Society for the Advancement of Science and at the Polytechnic Institute. To some extent the tensions are of the same kind that developed in engineering schools everywhere in the 19th century125 (and that are still present) about the role of theory for the practice of engineering: what level of generality is desirable in the education, so that future engineers may learn requisite principles without disappearing into an ivory tower. Hans Christian crossed swords with Grundtvig once again over the appropriate nature of the Polytechnic, and he needed to argue convincingly against the charge of irrelevance: It is easier to understand the most applicable of these doctrines than most think. When the sentences are only expressed in their natural simplicity and illustrated by experiment, the audience needs not so much a great collection of prior knowledge as rather sound judgment and attentiveness.126

and the charge of arrogance: The scientist has to be careful to not teach the citizen what the latter understands better: his own handicraft. […Their appropriate] relationship is a friendly, mutual communication and advice, not merely a one-sided teaching and prescribing by the scientist.127

With the changing nature of society, criticism of Hans Christian increased. Similar developments took place in France, where in the 1820s and 1830s a new generation 120

121 122 123 124 125

126 127

H. C. Ørsteds naturvidenskabelige Skrifter, edited by Kirstine Meyer, 1920 (n. 37), vol. II, 411–413, 497. Ibid. pp. 413–415 Ibid. p. 478 Ibid. p. 482 Ibid. p. 500 For example Karl-Heinz Manegold, Universität, technische Hochschule und Industrie; ein Beitrag zur Emanzipation der Technik im 19. Jahrhundert unter besonderer Berücksichtigung der Bestrebungen Felix Kleins, (Berlin: Duncker & Humblot, 1970). Kirstine Meyer, “Arbejdsliv” (n. 37), at XCIV–XCV. Kirstine Meyer, “Arbejdsliv” (n. 37), at CIII.

48

A. HESSENBRUCH

of physicists, such as Fourier, became adamantly positivist, meaning that they insisted on sticking with the observable and did not even address metaphysical questions.128 For example, when examining heat flow in a metal bar, one was to discuss only the temperature measurements that one could observe and the mathematical expressions that would summarize these measurements. One was not to consider the question of the nature of heat; that was metaphysics, not science. Hans Christian was criticized in just this vein,129 and the word Sværmeri (a state of dreaming, with overtones of passion and fanaticism) became of importance.130 Hans Christian repeatedly attempted to deflect the stigma attached to the word by defining it as not conforming to the ideals of transparency and accountability, but since he adhered closely to what contemporaries despised as metaphysical, the criticism stuck. In the 1830s, Hans Christian began to discuss the importance of politeness. He had now achieved power but public debate was growing and it was also increasingly critical. The assemblies of the estates was the most obvious forum for public debate, where an “appropriate democratic tone” had to be found, but the growing world of debate in the journals was also often ad hominem, hostile, and sarcastic. This was the period when Søren Kierkegaard sharpened his wit, and clearly Hans Christian did not approve. In obituaries of the old elite, especially Count Schimmelmann, he praised the “correct way of conducting differences of opinion in public.”131 Hans Christian emphasized the virtue of self-control, that he had found missing in Grundtvig (and that was an old mainstay of polite discourse, as discussed in the introduction). Control of one’s passions was generally speaking a virtue, but specifically it was crucial in science. One part of the dawning nationalist Danish discourse was the popular engagement with Norse antiquity. B. S. Ingemann had written the first blockbuster tales of valiant war heroes going berserk, and Hans Christian argued that the admiration of passions was not such a good idea. He praised the natural mild and peaceful Danish virtues, arguing also that not only passionate human beings can be geniuses.132

128

129

130

131

132

It is most interesting that Hans Christian conducted experiments with Fourier in Paris in 1823 and that their joint publication is in the style of Fourier. Hans Christian describes his prejudices against the French in “Ørsted (Hans Christian),” in Conversations-Lexicon (n. 2), at 524. He describes the much more sympathetic stance he developed to Parisian savants in a letter, quoted in extenso in Hauch (n. 2), pp. 153–155. For example, Repp criticized both the style of teaching and the lack of mathematical acuity in his 1844 textbook, Kirstine Meyer, “Arbejdsliv” (n. 37), at CLV–CLXI. Sværmeri corresponds quite well to romanticism with the pejorative overtones of uncontrolled passions. “Mindetale over det Kongelige danske Videnskabernes Selskabs Præses Hans Excellence Geheimestatsminister Ernst Heinrich Greve af Schimmelmann”, held 14 July 1831, published in Efterladte Skrifter (n. 2), vol. 6, pp. 50–70, esp. pp. 67–69. “Danskhed—en Tale” (Dansk Folkeblad, 1836), reprinted in Efterladte Skrifter (n. 2), vol. 7, pp. 39–58. Hans Christian had discussed the role of the genius in his 1807 history of chemistry but not in terms of control of passions. His 1807 discussion merely has the genius as ahead of the times and not essential; science progresses with a general momentum within which no individual is indispensable, “Betragtninger over Chemiens Historie, en Forelæsning,” Efterladte Skrifter (n. 2), vol. 5, pp. 1–33, at 25–27.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

49

8. BRANDED REACTIONARY Anders Sandøe’s role at the first assemblies of the Estates was considerable. Between 1830 and 1835 he worked hard to make the assemblies an advisory institution for the government, through which genuine two-way communication between government and general populace would take place.133 Against opposition within the central government he pushed through that the populace be widely represented, including copyhold peasants, civil servants, pastors, and academics (Videnskabeligheden, literally the scientificity).134 He also helped make sure that all legislation concerning citizens’ rights, taxation and other burdens (e.g. military service), would be brought to the assemblies for discussion before being passed into law.135 He enjoyed great popularity with the assemblies’ delegates, and, along with Professor Schouw, representative of the intelligentsia and a close friend of the Ørsteds, generally influenced the proceedings, including “giving the work at the assemblies an appropriate parliamentary form, which according to contemporary reports caused many of the generally un-experienced representatives difficulties.”136 Both Anders Sandøe and Hans Christian actively combated censorship. In the mid-1830s new more radically liberal opinions were gaining prominence, and the government kept a close watch on the press. In 1835, a royal statement containing the phrase “we alone know” what is best for the country backfired and became a rallying cry for opponents of absolutism until its demise in 1848.137 The brothers opposed the government’s intransigence, probably fearing the very upheavals that in fact did take place in 1848. They were cofounders of a Society for Free Printing, and both signed an 1835 petition against censorship.138 In the 1840s a new liberalism too radical for the Ørsteds’ taste gained ground and the narrower national Danish consciousness took shape, very much in opposition to German. Confrontations between Danish and German in SchleswigHolstein became a daily occurrence. The new national liberals claimed to speak for the peasantry, challenging the notion that the king was the safeguard of the interest of the broader population. Reference was made to remnants of feudal relations: copyholding, military service, and tax on peasants’ land only. In 1848 the absolutist government collapsed and was replaced by a democratic and constitutional monarchy. Of the 193 members of parliament 48 were to be appointed directly by the government, and the rest were to be elected. The franchise was given to all men aged 30 with an independent household not receiving any poor law subsidy. Censorship was abolished.139 The Ørsted brothers had grown up in a society with rigid class distinctions, where it was taken for granted that peasants, the largest part of the population, 133 134 135 136 137 138 139

Claus Bjørn, Fra reaktion til grundlov (n. 11), p. 198. Ibid. p. 189. Ibid. p. 191. Ibid. p. 198. Ibid. pp. 207–212. Kirstine Meyer, “Arbejdsliv” (n. 37), at CLIV. Claus Bjørn, Fra reaktion til grundlov (n. 11), pp. 338–339.

50

A. HESSENBRUCH

were vulgar and as such had to be treated differently, also in legal terms. Their philosophy of liberating the entire populace was radical: they insisted that there were no inherent differences between any human beings, globally, and that anyone could potentially rise to higher status in society and to a higher religious consciousness. By 1848, when Hans Christian was 71, this no longer was a radical and liberating political philosophy. Now, the Ørsteds’ conception of a hierarchy based on the one-dimensional scale of cultivation seemed elitist. The close friend of the Ørsteds and Professor at the University, J. F. Schouw expressed very well the outrage that the Ørsteds also felt at the emerging political ideas: To demand that the sciences should be popularized in totum is ignorant and unreasonable—it is even more so to charge the savant with aristocratic thinking for not complying with this demand. It is as if one were to demand of poets that they only wrote for the lowest common denominator; it is like criticizing Shakespeare, Goethe, Plato, Kant, and Fichte for having composed ideas that only a few in society can comprehend. Complete equality in the Republic of the Spirit is an even more moronic idea than complete equality of property in the civil world. Such equality can only exist where everyone’s an idiot.140

Everywhere around them, their world was crumbling. In his poetry written in 1848 Hans Christian describes the civil war and lack of peace everywhere “maybe necessary for the victory of reason but still scary,” and he retreated to the safe haven of the family. “Deeply we feel that the enemy is also our brother.”141 The benevolent paternalism of the state was being diminished: by 1850 the new national liberal party demanded that the government leave actual trade and industry to private forces and that its role be restricted to the provision of an infrastructure.142 Hans Christian’s benevolent paternalism as a provider of education/cultivation for the improvement of the entire populace was being criticized too. The teaching at the Polytechnic had been criticized for its lack of relevance to the demands of manufacture and industry all through its existence but it emerged openly into published journals in the late 1840s and the criticism was accepted by the central government in a prominent report published in 1850,143 when demands were voiced to move the Polytechnic from the Ministry of Culture to the Ministry of the Interior.144 Industrial production increased; the first Danish railroads for instance, were built in the 1840s. Technical education was now to produce skills applicable to engineering tasks, not a cultivating of morality through contemplation of the sublime.

140 141

142 143

144

Redact., “Om Naturvidenskabernes Fremstilling for Folket,” Dansk Ugeskrift, No. 15, 1832. Efterladte Skrifter (n. 2), vol. 4, pp. 17–18, 37. This seems to refer primarily to the uncivil tone in political debate, and perhaps specifically to Germans. The enemy of many was of course literally Hans Christian’s brother: Anders Sandøe held various ministerial posts between 1842 and 1853, including that of prime minister. His reactionary policies have meant that contemporaries attacked him viciously; and 20th-century secondary literature is much less fulsome with its praise of his legal than his political work. Michael F. Wagner, Det polytekniske gennembrud (n. 19), p. 387. Ole Jørgen Rawert, Danmarks industrielle Forhold indtil 1848, (København, 1850); also Michael F. Wagner, Det polytekniske gennembrud (n. 19), p. 391. Ibid. p. 382.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

51

The absolutist state within which the Ørsteds had successfully made their careers was dismantled. The franchise remained limited, to be sure, but it still represented governance by others than those who know best. Danish nationalism also grew, as did resentments against anything German. Anders Sandøe pleaded in vain with a Danish nationalist who provoked outrage in the Schleswig-Holstein assembly by demonstratively speaking Danish. Even their close friend, Professor Schouw, broke with the two brothers over the national issue. As a part of the rearguard action, Hans Christian argued that the Danish landscape resembled that of Northern Germany more than that of the rest of Scandinavia, the implication being that a putative national Danish nature would resemble that of Schleswig-Holstein.145 In 1850, the two brothers held high positions, but the world had changed and the next generation was dismissive or even hostile. Faced with the indictment of his life’s work, Hans Christian penned Ånden i Naturen, to explain to the uncomprehending world, just what the grander purpose was. Of course, by 1850, his discourse convinced nobody. It now fell foul of the charge of metaphysical thinking, the adjuring tone resembled the now less well-regarded preaching from the pulpit. Hans Christian seemed less of a mahatma and more of “romantic,” the next generation’s dismissive and pejorative term for the old and passé, and the very term that Hans Christian resented because it connoted a lack of control of one’s passions.

9. CONCLUSION AND DISCUSSION 9.1. Freedom, morality, trust, and tact The topics of morality, trustworthiness, and individual freedom are of general importance. In the patrician culture of early modern natural philosophy, trust and honour developed through face-to-face interactions that were immensely dependent on the apparently trivial and superficial marks of comportment and behaviour. The society of orders bred, according to Norbert Elias, an “extraordinarily sensitive feeling for the status and importance that should be attributed to a person in society on the basis of his bearing, speech, manner or appearance.” Elias even argues that this culture of restraint and self-control accompanied the emergence of the modern polity.146 This issue continued to matter during the Ørsteds’ lifetime. The report of a famous commission to examine mesmerism in Paris in the 1780s is telling with regard to this issue because it denounced situations subverting the necessary separation of reason from bodily passion, and as Alison Winter has shown the issue continued to plague the sciences in the 19th century too.147

145

146 147

Hans Vammen, “‘Schouw er velsignet’—En professor og hans guldaldernetværk”, in Krydsfelt— Ånd og natur i Guldalderen, edited by Mogens Bencard, (København: Gyldendal, 2000), pp. 246–257. Norbert Elias, The Court Society, (Oxford: Oxford University Press, 1983), p. 55. Report of Dr. Benjamin Franklin, and other Commissioners, charged by the King of France, with the examination of the Animal Magnetism, as now practised at Paris, London, 1785; Alison Winter, Mesmerized: Powers of Mind in Victorian Britain, (Chicago, IL: The University of Chicago Press, 1998).

52

A. HESSENBRUCH

These issues are clearly present in Hans Christian’s career and tied up with his way of doing science: self-control, tact, cultivation of morality through the encounter with the encryption of God in Nature, and the authority of the cultivated. He was a deist natural philosopher living through a transition from a purely religious society to a semi-religious, bourgeois one. He was not in favor of a wholehearted transition to secularism but rather wanted to keep all aspects of religion within the new framework of a new civil order. His life spanned a development of Danish society from an absolutist and feudal state with a minimal public sphere to a constitutional monarchy, a parliamentary democracy, and an economic sphere based primarily on contractual relationships. From the 1790s onwards, the expansion of the civil service led to its increased ability to raise taxes. Evermore professional surveyors measured the newly privatized and enclosed land, gaining in authority and visibility. They partook in measurements for the purpose of indirect taxation at town gates, harbors, mills, public houses, and turnpikes. In other words, authorized experts using authorized weights and measures were penetrating into the nooks and crannies of society (Fig. 1). Now, a most important aspect of the work of these professional measurers was that they were to be trustworthy; dexterity was not the issue. When measuring barrels of grain for the purposes of taxation, the most important issue was the height from which one poured the grain. It required no great skill to pour from a specified height, but it did require tact, a cultivated state of being. It is this issue that weighed upon Hans Christian’s mind, not the concern of teaching engineers

Fig. 1. Trust, tact, morality and authority in action: a Copenhagen town gate in the early 1840s. The invigilating authority of the state, exemplified by an array of soldiers, is on the right. The building to the left is a weighbridge, and peasant carts on their way to market pass through the bottleneck paying a tax.

THE MAKING OF A DANISH KANTIAN: SCIENCE AND THE NEW CIVIL SOCIETY

53

manual skills. Trust mattered for the new civil order; manual skills did not. That is why he used the word “tact” to describe what is required of the experimenter; it refers to social responsibility, not to dexterity. And trust was the central issue for Anders Sandøe too, the Pollux to Hans Christian’s Castor. For example, when the Danish state had been forced to declare bankruptcy in 1812 (Copenhagen had reluctantly given up the lucrative trade under a neutral flag of the late 18th century, backed the wrong horse in European politics, France, and in 1807 the British fleet bombarded Copenhagen and took with it all of the Danish fleet it hadn’t sunk) Anders Sandøe became intimately involved with the resultant currency reform. He was guided by the principle that contracts ought to be honored as far as possible. This involved sacrifices on the part of debtors and creditors, and he sought to find an equitable solution in a staggered debt write-off depending on the time the contract had been entered into. In other words, the currency reform was intended to minimize damage to contractual arrangements. Contracts depend upon disseminated trust, a kind of trust in the system that arrangements will be adhered to in the future. Thus the maintenance of contractual arrangements is co-terminous with maintenance of disseminated trust. Another aspect of the same issue is that individuals be held responsible for their actions, and Anders Sandøe repeatedly insisted on culpability of the individual citizen. In his jobs as judge and civil servant he continually encountered the issue of intentionality, and he always aimed to clarify the notion of culpability with the aim of facilitating the smooth working of contractual arrangements, that is to support a disseminated trust wherever possible.148 He was guided by the same principle when dealing with witnessing and proof149: both depended upon trust and trustworthiness. Thus the concepts of trust, tact, morality, and authority were central to Danish society and to the Ørsteds. Very late in his life, at the age of 73, Hans Christian penned The Spirit in Nature, a collection of articles explaining his philosophy to a by then uncomprehending world. An 1851 unfinished 60-page essay, The Road from Nature to God, rehearses the same argument but is even more explicit about the political nature of the two Ørsted brothers’ perspective. It ties together all the themes: the public character of knowledge, the sublime encrypted in nature, the focus on morality instead of free will, expertise and trust, transparency, self-discipline, and hierarchy defined by levels of cultivation. My hope is that by now the following summary of this essay will evoke the Danish context just as much as Kant’s writings. The argument starts from a consideration of the heliocentric system. All experts agree that it is correct and yet it contradicts the immediate evidence of the 148

149

Ditlev Tamm, Fra “lovkyndighed” til “retsvidenskab” (n. 81), p. 25; Troels G. Jørgensen, De Ørstedske Straffelove (Copenhagen: Gyldendalske Boghandel Nordisk Forlag, 1948), pp. 63–66. Much of his activity relates to this issue: ibid.: The nature of denial under oath (42), burden of proof (56), sanity (63), denial under oath may not count as evidence (73), juramentum ignorantiæ (76), circ*mstantial evidence (81), presumption of reality of public documents (86), acceptance of affidavit (119), punishment for lying under oath (121), trustworthiness of police officers (123), proof of lie under oath (131), swearing in a Jew (134), abuse of force during interrogation (142), proof of causal occurrence in murder cases (143), sanity in murder case (145, i mordsag 162), punishment for calling unnecessary witness (159).

54

A. HESSENBRUCH

senses. Any layperson doubting it may either learn to understand the argument or simply trust the expert. The heliocentric system is based on fundamental laws, such as the law of gravity. The predictions of astronomical events known from the almanacs provide striking evidence that this law is correct. Such laws govern the entire world, including the animal world. The earth has developed in the long term through the impact of forces that have brought geological changes and simultaneously corresponding changes to the animal population. To understand these encrypted laws of the whole world, from the smallest creature to the entire universe, is to reach the sublime: it amounts to understanding God. Now, within this world, the only creature with freedom is the human being. The strivings of human beings have a unifying and a dispersing effect. Sexuality, filial and parental love, sociability, and the need for communication, and the need for mutual help unify. Desire for the same goods, limited nourishment in one place, and the attraction of novelty cause dispersion. Wise men constitute a special unifying force through their ability to iron out disputes, give advice for and prescribe peaceful coexistence, to heal, and to improve tools. Truths have to be openly available; secrecy leads to abuse of power. The more societies have evolved, the more such advice and prescriptions have become laws, and those laws that have been perceived as genuinely protective, have been accepted. The human being chooses, out of his free will, to obey. Duty develops. There has been a historical development of education/ civilization and former centers include Egypt, India, Persia, China, and Greece. The development tends toward a global higher civilization without ever reaching perfection. All religions contribute, but Christianity is special, because it foregrounds love of the spiritual life. Although many human souls are closed to the blessed divine truth due to vulgarity, passion, prejudice, and misleading scientific opinions, it is possible to open them. The will to live an orderly life (in accordance with Divine reason) constitutes virtue, and virtue develops and defines duties that all contribute to order and reason. An existence catering solely to one’s pleasures will remain empty, whereas a life working for higher order gives meaning. This virtue is much more important than the mere existence of a free will. A higher national order can not be achieved through the imposition of a speculatively rational legislation. Rather, piecemeal improvements in legislation have to adapt to the already existing social realities. The more civilized should improve upon the less so. Egoism is not to be rejected out of hand, merely to be cultivated. Most virtues are to be found between extremes: independence is between stubbornness and dependence, courage between recklessness and cowardliness, noble pride between haughtiness and self-contempt, humility between pride and self-humiliation, decisiveness between obstinacy and fickleness.150 Massachusetts Institute of Technology

ACKNOWLEDGMENTS I am very grateful to the many helpful suggestions from Bob Brain, Dan Ch. Christensen, Anja Skaar Jacobsen, Ole Knudsen, and Jeff Horn. 150

“Veien fra Naturen til Gud,” vol. 3 of Efterladte Skrifter (n. 2).

PHRENOLOGY AND DANISH ROMANTICISM ANJA SKAAR JACOBSEN

The appearance of phrenology in Copenhagen in the beginning of the 19th century is a fascinating story because it involved many prominent figures in medical, scientific, and intellectual circles.1 Several factors shaped its first reception here, for instance the general positivistic turn of both science and medicine from the 1820s onwards.2 However, what foremost coloured the discussions of phrenology in Copenhagen was the philosophical and intellectual climate at the time in terms of Romanticism. The present paper illuminates and discusses the reception of phrenology in Copenhagen on this background. Hans Christian Ørsted (1777–1851) was not directly involved in this process, but occasionally he expressed his thoughts on the matter. Moreover, as his scientific career matured, his position in Danish society consolidated and he came to constitute the centre per se of the literary and scientific elite in Copenhagen around whom the philosophical discussions of phrenology evolved.

1. INTRODUCTION When young Danish natural philosophers or medical men went abroad in the beginning of the 19th century in order to widen their knowledge of their respective fields at the European scientific and medical centres, many of them either heard about Franz Joseph Gall (1758–1828) or met him in person, attended his phrenological lectures while in Paris or in other European cities, or they met Gall’s former assistant Johann Gaspar Spurzheim (1776–1832) in London.3

1

2

3

The name phrenology was used for the first time about Gall’s doctrine in 1815 and came into general use in the 1820s. Although Gall did not recognise this name for his Schädellehre, the word phrenology is used in the following even in relation to Gall. Patricia S. Noel and Eric T. Carlson, “Origins of the word ‘Phrenology,’” American Journal of Psychiatry, 127: 5, (1970), pp. 154–157. I further denote authors of Naturphilosophie, such as Schelling and Steffens, Naturphilosophen while adherents of this philosophy and members of the Romantic movement generally I call Romantics. This perspective can be found treated in A. S. Jacobsen, “Carl Ottos forbryderhoveder—Frenologi og det intellektuelle miljø i København i første halvdel af 1800-tallet,” Bibliotek for Læger, to appear in 2004. Letter from Daniel Frederik Eschricht to his mother, Paris 8 August, 1824, the Royal Library, Copenhagen, NKS 3100 4°. Jens Veibel Neergaard, Om Phrenologien eller den saakaldte Gallske Hjerne- og HjerneOrgan-Lære (Copenhagen: den Wahlske Boghandling, 1827), p. 45. Mathilde Ørsted, ed., Breve fra og til Hans Christian Ørsted, 2 vols. (Copenhagen: Lind, 1870), vol. 1, pp. 45, 58–61. Carl Otto, Phrænologien eller Galls og Spurzheims Hjerne- og Organlære i fuldstændig Oversigt og i sine senere Fremskridt med Bidrag til dens nøiere Kundskab og Stadsfæstelse (Copenhagen: Brummer, 1825), p. iv.

55 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 55–74. © 2007 Springer.

56

A. S. JACOBSEN

In H. C. Ørsted’s travel diary from his journey through Europe 1801–1804 it appears that Gall’s theory was widely recognised at that time and apparently also by Ørsted himself. After he was shown a wax figure of Immanuel Kant (1724–1804) in Berlin 28 March, 1802, he wrote in his diary: “I have never seen the perspicacity organs, in correspondence with Gall’s [theory], as articulated as in Kant.”4 And upon a veritable quarrel about the justification of Naturphilosophie with the chemist and opponent of Naturphilosophie, Alexander Nicolaus von Scherer (1771–1824), the latter concluded by the remark: “Perhaps it is my fault; Gall says I do not possess the metaphysical organ,” and Ørsted answered “Yes […] in that Gall has not been mistaken.”5 Meanwhile, as other devoted Romantics, Ørsted was sceptical about Gall’s emphasis of the independent development of individual soul faculties. He discussed Gall’s Schädellehre with Johann Gottlieb Fichte (1762–1814), who opposed the theory but still admitted that Gall was right when he stated—after he had examined a bust of Fichte’s head—that he, Fichte, possessed a great sense of locality. However, Fichte and Ørsted agreed that the faculties of soul functioned as part of a whole and could not be developed individually the way Gall proposed. This attitude towards Gall’s Schädellehre reflects the general tendency among Danish (and German) Romantics. Many years later Ørsted stated that the origin of mental life was not something material, but was simply spiritual forces, which, apart from the fact that they were opposite and capable of uniting and leading to equilibrium, he did not specify any further.6 Although Gall’s Schädellehre was mostly introduced into medical circles through medical journals, it was not a new kind of alternative medical cure or health treatment like for instance mesmerism. Gall and Spurzheim announced phrenology as a science of man’s social behaviour. The basic principles of Gall’s phrenology were (1) that the brain was the physical embodiment of the mind and soul, (2) that the brain was made up of several (27) separate organs, i.e. placed in different and independent parts of the brain, each corresponding to distinct mental faculties, (3) that the relative size of a particular organ was a measure of the power of the associated faculty, (4) that the mental faculties were innate in both animals and humans, and—for practical reasons the most important of all— (5) that there was correspondence between the contour of the brain and the outer surface of the skull. On this background it was assumed that the relative size of the organs and consequently talents and propensities could be deduced from a craniological examination of the “bumps” on a person’s skull.7 The novelty of phrenology was the assumption that the mind operated through the cerebral organs. Not only was the brain assumed to be the organ of the soul and mind, Gall assumed no less than 27 independent organs related to the faculties 4 5 6

7

Ørsted op.cit. (3), vol. 1, pp. 45, 60–61. Ibid. pp. 58–59, see also p. 124. H. C. Ørsted, “Det aandelige Liv,” in Samlede og efteladte Skrifter af H. C. Ørsted, vol. 3 (Copenhagen: A. F. Høst, 1851), pp. 38–64, on p. 42. Gall, Franz Joseph (1979[1798]), [Letter to Joseph Friedrich Freiherrn von Retzer] in Erna Lesky, ed., Franz Joseph Gall 1758–1828: Naturforscher und Anthropologe; ausgewählte Texte, Hubers Klassiker der Medizin und der Naturwissenschaften, vol. 15 (Stuttgart and Wien: Hans Huber Bern) pp. 47–59, on pp. 48–53.

PHRENOLOGY AND DANISH ROMANTICISM

57

of mind located in the cerebral cortex. Thus he related specific parts of the brain to the different functions of the mind, soul, and psyche. New was also the apparent empirical method of demonstrating the correlation between the structure and the functions of the brain and human behaviour in terms of cranioscopy. Despite their great reputation as anatomists, Gall and Spurzheim did not base their phrenological theory on dissection and other anatomical investigations with regard to localising the cerebral organs. Instead they used the technique of observing external cranial contours on the living as well as the dead, humans as well as animals. They claimed that anatomists could never discover the function of the brain, or the tongue for that matter, by dissecting it. According to Gall and Spurzheim, neither the anatomist who examined the body nor the philosopher who investigated the mind would gain knowledge of the mind–body relationship; this was up to the phrenologists.8

2. PHRENOLOGY VERSUS NATURPHILOSOPHIE The reception of phrenology in Germany quickly developed into a dispute between Gall and the dominant Naturphilosophen who rejected nearly all of Gall’s basic principles. Friedrich Wilhelm Joseph Schelling (1775–1854) published a short, sceptical comment about Gall’s doctrine in the Morgenblatt in 1807, and in his physiology lectures in 1805 Henrich Steffens (1776–1845) clearly stressed some of the points where the Naturphilosophen and Gall disagreed.9 Steffens began by clarifying the background of naturphilosophisch thinking in the following way: It is the principal clause of Naturphilosophie to view things […] only in the sense and absolute harmony of the whole, not to separate what Nature has not separated; […T]he reality of functions is not situated in one or the other factor but in the unity of both.10

On the basis of this statement Steffens rejected Gall’s organology and argued that the brain had no reality in itself, i.e. separated from the body. Every single part of an organism is not an autonomous entity, but rather a part of a coherent whole, which in turn determines its parts.11 In other words, the brain could only exist in unity with all organs of the body. Therefore it was a misunderstanding to look for the organ of the soul and especially to look for a particular seat of it in the brain, or to fragment it according to several organs in the brain, as Gall did. Steffens thought the whole body was the organ of the soul and that thinking was an ability of the whole body, not particularly the brain. He concluded that Naturphilosophie and phrenology could never be united; it was a choice between 8

9

10 11

Martin Staum, 1995: “Physiognomy and Phrenology at the Paris Athénée,” Journal of the History of Ideas, 56(3), (1995), 443–462, pp. 443–462, on pp. 449–450. Schelling, F. W. J., “Einiges über die Schädellehre,” Morgenblatt, 1807, No. 74, republished in Sämtliche Werke, K. F. A. Schelling, ed., vol. VII, Abt. 1 (Augsburg: Cottascher Verlag, 1860), pp. 542–543. Henrich Steffens, Drei Vorlesungen des Herrn Prof. Steffens zu Halle über Hrn. D. Gall’s Organenlehre (Halle: Im Verlags der N. Soc. Buch- und Kunsthandlung, 1805). Ibid. p. 18. Ibid. p. 33.

58

A. S. JACOBSEN

accepting Naturphilosophie and denying phrenology’s cerebral organs or accepting phrenology and denying freedom and with it morality.12 On the methodological level, the Naturphilosophen criticised Gall for having excluded metaphysics; his procedure was unphilosophical. In other words they criticised his purely empirical approach. Indeed Gall considered his own “inductive” method completely free of speculations and hypotheses and claimed to employ only empirical investigations and cold comparative observation. He on his behalf regarded Naturphilosophie as pure mysticism and vision with its emphasis on “intellectual intuition” and “speculative physics.”13 The Naturphilosophen did not question Gall’s anatomical skills, in fact few did, but whereas Gall sought to relate physiology and anatomy to psychology, the Naturphilosophen generally doubted the relevance of anatomical discoveries for the knowledge of human spiritual life, in a way implying maintenance of dualism between the two domains. In a time when psychology approached biological methods distancing itself from philosophy, Schelling maintained that spiritual life was not an object to be investigated by physiology or anatomy. Anatomists and physiologists should continue to do empirical investigations while the interpretations of their results were the philosopher’s task. The Naturphilosophen could also not accept the materialistic implications of phrenology, and that it led to materialism was probably the most common accusation against phrenology. Despite deviancies on the methodological level there were apparent philosophical correspondences between phrenologists and Naturphilosophen. For example, both were anti-dualists rejecting a soul independent of body, but phrenologists fragmented the soul into various functions while Naturphilosophen maintained an indivisible soul. The main factor, displaying Schelling’s rejection of dualism both ontologically and epistemologically, was his philosophy of the original identity between nature and soul, i.e. the assumption that the same intelligence exists in nature and in our thinking. As Schelling expressed it: “Nature should be Mind made visible, Mind the invisible Nature,”14 and generally Naturphilosophen claimed identity of soul and brain interpreting them as polarities i.e. antagonistic aspects of “the Absolute.” Schelling agreed with traditional mind-body dualists that the essence of the self was spirit and the essence of nature was matter, only he assumed that the essence of matter as well as life was force or activity—attraction and repulsion. Thus he found a common ground of nature and the self in activity and could maintain that mind was nature. Nature was viewed as a living organism, as natura naturans, i.e. creative, undetermined, free, productive, and infinite activity. It was supposed that nature contains the organic within itself, inorganic matter being matter that had ceased to live. Thus the inorganic world was subordinated the organic and not vice versa. Life was the bearer of nature 12 13

14

Ibid. p. 29. John van Wyhe, “The Authority of human nature: the Schädellehre of Franz Joseph Gall,” BJHS, 35 (2002), 17–42, p. 38. Michael Hagner, “The Soul and the Brain between Anatomy and Naturphilosophie in the Early Nineteenth Century,” Medical History, 36 (1992), 1–33, p. 13. F. W. J. Schelling, Ideas for a Philosophy of Nature [1797/1803], tr. E. E. Harris and P. Heath (Cambridge: Cambridge University Press, 1988), p. 42.

PHRENOLOGY AND DANISH ROMANTICISM

59

and also of mind since mind was nature. Therefore, like nature in general, mind could not be originally material object, such as suggested by phrenology for instance, but became object only through its own acting. Mind was regarded as a productive faculty.15 When phrenologists announced that human affairs were ruled by the universal laws of nature, and that abilities or propensities of individuals were innate, i.e. part of the basic underlying order of Nature, this seems not completely at odds with the basic principles of Schelling’s philosophy of identity. Both Naturphilosophen and phrenologists supposed a Spinozan conception of mind and matter as two different attributes of the same substance. However, the Naturphilosophen based it on the assumption that both nature and mind was spiritual activity whereas the phrenologists reduced mind to physical organs, i.e. based both nature and mind on a material origin. On the methodological level both parties insisted on the unity of man and nature; the study of subjective mind should be pursued according to the same methods as employed in the science of body. However, Naturphilosophen claimed the unifying method was speculation while Gall claimed it was the method of induction.

3. GALL IN DENMARK Gall visited Copenhagen and other Danish cities (Odense and Kiel) in the fall of 1805, as part of his European lecture tour. Gall’s visit was extremely well documented in contemporary newspapers compared to other “scientific” events in Denmark. As in other European cities it was a sensation and subject of general discussion. Gall was received with the greatest respect, enormous enthusiasm, and interest by both the city of Copenhagen and the university, and the large university auditory was immediately placed at his disposal. He gave several courses during the one and a half month duration of his visit. The crown prince Christian (later King Christian VIII), the councillor of the university, and several members of the government as well as several Swedish scientists, attended his lectures. It was possible to buy a silver medal with Gall’s portrait on the one side and a human skull on the other as well as plaster casts of skulls with drawings of Gall’s organs on them. Gall visited the Danish prisons only to confirm his cranioscopic predictions when he estimated the heads of the criminals.16 15

16

For accounts of Schelling’s Naturphilosophie and system of identity see Frederick C. Beiser, German Idealism—The struggle against subjectivism 1781–1801 (Cambridge, Massachusetts: Harvard University Press, 2002). Elke Hahn, “The Philosophy of Living Things: Schelling’s Naturphilosophie as a Transition to the Philosophy of Identity,” in William R. Woodward and Robert S. Cohen, eds., World Views and Scientific Discipline Formation, BSPS, vol. 134, (London, Boston, Dordrecht: Kluwer Academic Publishers, 1991), pp. 339–350. Bernhard Rang, Identität und Indifferenz. Eine Untersuchung zu Schellings Identitätsphilosophie, Philosophische Abhandlungen, vol. 78 (Frankfurt am Main: Vittorio Klostermann, 2000). Carl Otto, “Phrenologiens Historie i Danmark,” Tidsskrift for Phrenologien, 1 (1827), pp. 353– 372. E. Snorrasson, “The Danish Physician Carl Otto (1795–1879) and Phrenology,” in Wien und die Weltmedizin, Lesky Erna, ed. (Wien, Köln, Graz: Verlag Hermann Böhlaus Nachf., 1974) pp. 146–158, on pp. 151–152 Villads Christensen, “Dr. Galls Ophold i København 1805,” Historiske Meddelelser om København, 8 (1921–22), pp. 217–225.

60

A. S. JACOBSEN

During Gall’s visit mainly the philosophical implications of phrenology were seriously questioned and were not tolerated in respectable circles. Thus Gall visited the house of the Prime Minister Count Heinrich Ernst Schimmelmann (1747–1831), who had also invited prominent Danish physicians and natural philosophers to meet Gall (we are not told who). They praised Gall for his anatomical skill but criticised him for his philosophical inferences.17 Numerous publications on phrenology appeared in Denmark in immediate continuation of Gall’s visit.18 In the period 1810–1819 little was written about phrenology, but it was introduced in lectures on physiology and psychology at the Copenhagen University (and at the University of Kiel). Johann Daniel Herholdt (1764–1836) introduced phrenology in his lectures on physiology, but he rejected it completely. He had thoroughly and critically discussed phrenology from a physiological perspective already in 1803. By contrast, the professor of philosophy Frederik Christian Sibbern (1785–1872), who also introduced phrenology in his lectures on empirical psychology, supported it.19

4. OTTO’S PHRENOLOGICAL CAMPAIGN The history of phrenology entered upon a new phase when Spurzheim left Gall and Paris in 1814 and began his own popularising lecture activity in Britain. Spurzheim’s version of phrenology varied from Galls original ideas, for instance by stating the existence of 35 cerebral organs rather than 27 and by omitting all “evil” organs. Spurzheim was more optimistic than Gall with regard to the application of phrenology in education and in the treatment of insane people. He saw phrenology as a means to improve on mankind and his phrenological classification soon turned more popular than Gall’s.20 Spurzheim’s activity led to an explosion in the numbers of adherents of phrenology, active phrenologists, phrenological societies, and journals, in other words the institutionalisation of phrenology in Britain. The repercussions of this movement also reached Copenhagen through the physiologist and phrenologist Carl Otto (1795–1879) who got inspired from his travels in Britain to start his own phrenological campaign in Copenhagen in the beginning of the 1820s. 17

18 19

20

Louis Bobé, Statsminister Greve Heinrich Ernst Schimmelmann (Copenhagen: printed by Nielsen & Lydicke, 1902), p. c. For a list of publications see Otto op. cit. (16), pp. 359–363 Johann Daniel Herholdt, “Nogle Anmærkninger over Doctor Galls Lære om Hiernens Forretning,” Skandinavisk Museum ved et Selskab, 1 (1803), pp. 215–246. Otto op. cit. (16), p. 368. F. C. Sibbern, Mennskets aandelige Natur og Væsen. Et Udkast til en Psychologie, 2 vols., vol. 1 (Copenhagen: Gyldendal, 1819), p. 48. Gall made a lecture course in Kiel returning from Copenhagen, which was arranged by the medical professors Christoph Heinrich Pfaff (1773–1852) and Johann Leonard Fischer (1760–1833). Among the audience to his lectures in Kiel were several nobles (the Rewentlows) and people in the administration of Schleswig-Holstein. Apparently the philosophical implications of phrenology were not discussed much in Kiel. Soon after Gall’s visit Pfaff introduced “Gall’s most important discoveries” into his lectures on physiology. Peter-Christian Wegner, “Franz Joseph Gall in Schleswig-Holstein,” Zeitschrift der Gesellschaft für Schleswig-Holsteinische Geschichte, III (1986), pp. 119–141. Christoph Heinrich Pfaff, Lebenserinnerungen (Kiel: Schwers’sche Buchhandlung, 1854), pp. 275–276 Owsei Temkin, “Gall and the Phrenological Movement,” Bulletin of the History of Medicine, XXI(3) (1947), pp. 275–321, on p. 308.

PHRENOLOGY AND DANISH ROMANTICISM

61

After returning to Copenhagen, Otto gave a private course on physiology in 1822–1825, in which he also presented phrenology. In 1825–1826 he gave a pure phrenology course in which he illustrated the principles of phrenology by means of plenty of images, skulls, and busts. The audience for this course amounted to 60 people, according to Otto. He followed up on his lectures by publishing his book Phrænologien eller Galls og Spurzheims Hjerne- og Organlære (Phrenology or Gall’s and Spurzheim’s cerebral and organ doctrine) as well as articles on phrenology in his own partly popular, medical journal Nye Hygæa (New Hygeia). According to himself, his book was well received and was quickly sold out, but his colleagues “turned up their noses at it.” Thus the leading medical journal Bibliothek for Læger (Library for physicians) reviewed the Nye Hygæa and particularly the articles on phrenology in a derisive tone.21 In 1825 Otto was employed as physician at the prison in Copenhagen. Otto was quite happy about this employment since it enabled him to make phrenological studies of the criminals. In his memoirs he tells the following anecdotes: at the death of “notable criminals at the institution […] always before their funeral [… I let] their head be cut off […] in order to keep them in my phrenological museum.” Every time a murderer was beheaded in Copenhagen or in the provinces, Otto attempted to get hold of the heads, often with success.22 By means of bribery he made the executioner in Copenhagen “his friend.” In 1838 a young man, Petri Worm (1814–1838), who was out of a well-off family strangled a “wellknown eccentric” newspaper publisher and the suspect was sentenced to death by beheading. The murder and the atypical background of the murderer were subject of general conversation in the capital. Apparently he had made confessions about his deed, he was good-natured despite this deed and even had a poetic talent. These circ*mstances made Otto quite anxious to get hold of Worm’s chopped-off head in order to study it phrenologically. He attempted to bribe the executioner as usual, but Otto’s pursuits as a phrenologist had become well known among the prisoners and Worm attempted to prevent his head from ending up in Otto’s phrenological museum. Worm explicitly made the police and bailiff responsible that his head would follow him into the coffin, which was in fact fulfilled. Meanwhile Otto was prepared to stick at nothing to get hold of Worm’s head. He had the head removed from the grave making a stir among the lower classes in the capital since they had come to regard Worm almost as a martyr. However, the police director knew about Otto’s phrenological pursuits and did not follow the case any further. Whether anecdotes like this were the motive for one of the prisoners’ 21

22

Snorrasson op. cit. (16), p. 150. Otto op. cit. (3), pp. iv–v. Carl Otto, Af mit Liv, min Tid og min Kreds (Copenhagen: L. A. Jørgensens Forlag, 1879), p. 200. Anonymous, “Ny Hygæa, udgivet af C. Otto,” Bibliothek for Læger, 5 (1825), pp. 335–343. Otto 1879 op. cit. (21), pp. 83, 205. When he resigned as professor at the university in 1862, Otto donated his collection of skulls as well as plaster casts of facemasks to the “physiological collection” at the Copenhagen University, where they were “beautifully exhibited,” according to himself: ibid. p. 209. C. Goos, Aarbog for Kjøbenhavns Universitet, den polytekniske Læreanstalt og Kommunitetet, indeholdende Meddelelser for de akademiske Aar 1864–71, vol. 1 (Copenhagen: Gyldendal, 1887), p. 452. Snorrasson op. cit. (16), pp. 148, 150. Most likely the phrenological skulls still existing at the Anthropological Laboratory at the Copenhagen University and at the Medical Museion originate from Otto’s collection (some of them shown at the images).

62

A. S. JACOBSEN

attempted murder on Otto in 1843 is not clear, but Otto asked for his resignation after this incidence.23 Upon his successful lectures and the reception of his book on phrenology in 1825, Otto published a journal Tidsskrift for Phrenologien (Journal of Phrenology) in 1827–1828. Only Otto contributed to the journal by “some of the best productions of [his] mind.” The publication ceased in 1828 and the last contribution to the journal happened to be a translation of François Joseph Victor Broussais’ (1772–1838) speech at Gall’s funeral the same year. Otto now devoted his time to editing the medical journal Bibliothek for Læger, and from 1832 to his position as professor of pharmacology and forensic medicine at the university.24 Carl Otto’s energetic appearance and great endurance in favour of phrenology in the 1820s produced a stir in medical as well as in the academic circles in Copenhagen. In the following I focus on the reaction in academic circles. The discussions of phrenology following Otto’s campaign coincided with a philosophical dispute about free will which bordered closely on the issues raised by phrenology and which reflects the intellectual climate in Copenhagen at the time.

5. THE FREE WILL DISPUTE AND THE INTELLECTUAL ATMOSPHERE IN COPENHAGEN Otto’s predecessor as professor of pharmacology and forensic medicine as well as in the position as physician responsible for the prison in Copenhagen, Frantz Gotthard Howitz (1789–1826), published in 1824 a treatise Om Afsindighed og Tilregnelse (About insanity and sanity) in which he claimed that criminals were not necessarily morally irresponsible but were determined in their act due to the structure of society and their own mental organisation. Phrenology was not a central issue in his treatise, but Howitz drew on Spurzheim’s phrenological ideas with regard to how the degree of reason in a person was reflected on the forehead of that person. Howitz took for granted that moral nature and will was determined, as was the intellect, by the organisation and development of the brain and on innate dispositions. He was convinced that every psychic process had its basis in a materially defined state of the human organs especially the brain. Therefore to talk about the freedom of will was to Howitz an illusion. He supported his arguments by David Hume’s (1711–1776) deterministic philosophy and criticised Kant’s views of the matter.25 23

24

25

Otto 1879 op. cit. (21), pp. 89, 205–207. Sidsel Eriksen, “Mellem medicin og moral. Eller hvorfor frenologen Carl Otto og rationalisten Carl Holger Visby begge var interesserede i den dødsdømte Petri Worms hoved,” in Historie og Historiografi. Festskrift til Inga Floto, Carsten Due-Nilsen, ed. (Copenhagen: Den danske historiske Forening, 2002), pp. 58–78. Otto 1879 op. cit (21), pp. 166, 201. See also letter from Otto to F. C. Sibbern 1849, Copenhagen: The Royal Library, Add. 1040 4°. Franz Gothard Howitz, “Om Afsindighed og Tilregnelse,” Juridisk Tidsskrift, Anders Sandøe Ørsted, ed. 8(1) (1824), pp. 1–117, on pp. 41–42. Determinismen eller Hume imod Kant-Et philosophisk Forsvar for Afhandlingen om Afsindighed og Tilregnelse (Copenhagen: Seidelin, 1824), p. v. J. P. Mynster, “Determinismen eller Hume mod Kant. Et philosophisk Forsvar for Afhandlingen om Afsindighed og Tilregnelse af Dr. Med. Franz Gothard Howiz,” Dansk Litteratur-Tidende, No. 1–4 (1825) pp. 1–48, 53–64, on pp. 4–6.

PHRENOLOGY AND DANISH ROMANTICISM

63

The treatise raised a philosophical dispute involving such prominent heads as H. C. Ørsted’s brother, the influential civil servant and later Prime Minister, Anders Sandøe Ørsted (1778–1860), the theologian and later bishop Jacob Peter Mynster (1775–1854), the aestheticist Johann Ludvig Heiberg (1791–1860), and the professor of philosophy F. C. Sibbern.26 Howitz disagreed with the prevailing Romantic view since he considered the origin of madness to be not in the immaterial soul but in actual physical disturbances in the brain. Howitz’ critics did not care much about his ideas of mental deviancies and their judgement in criminal law. In fact they even agreed with him to a certain extent. Thus, H. C. Ørsted later “dare[d] to claim that almost all of them [the criminals] have a certain crookedness in their mental predispositions.”27 Rather it was Howitz’ philosophical engagement which aroused the debate. H. C. Ørsted, who discussed Howitz with J. L. Heiberg in 1825, reveals the real issue at stake in the dispute from the following statement: The reason I am angry at Howitz is not because of his point of view, but because he with so much haughtiness and incorrect bragging of having read books of which he understands nothing, undertakes a hectoring tone towards people with superior knowledge.28

This statement is quite characteristic of the intellectual atmosphere in Copenhagen during these years. The common ideological background of the academic elite was Romanticism inherited from Germany. Steffens brought Schelling’s ideas to Copenhagen in a personally modified form through a series of lectures in 1802–1803 held at Ehler’s College. These lectures initiated the Romantic era here: a Schellingian storm now raged in intellectual life in Copenhagen involving a close and strong exchange among philosophical ideas, literature, science, art, politics, theology, social theory, law, etc.29 Politically the Romantics were liberal, but in contrast to the following generation of ultra or national liberals, the Romantics were more moderate liberals who co-worked well with the administration and the absolute monarchy.30 26

27

28 29

30

The Howitz dispute has been extensively treated in the history of philosophy in Denmark. See for example Knud Waaben, Retspsykiatri og strafferet i historiens lys (Copenhagen: Janssen-Cilag A/S, 1997). Carl Henrik Koch, “Ørsted og striden om viljens frihed,” in Anders Sandøe Ørsted 1778– 1978, Ditlev Tamm, ed. (Copenhagen: Juristforbundets Forlag, 1980), pp. 87–121. Oluf Thomsen, F. G. Howitz og hans Strid om “Villiens Frihed” (Copenhagen: Levin & Munksgaard, 1924). H. C. Ørsted, “Bemærkning om Forbrydernes Sjæleevner,” in Samlede og efterladte Skrifter af H. C. Ørsted, vol. 9 (Copenhagen: Høst, 1852), pp. 100–101, on p. 101. Although Ørsted then recognised the existence of individual instincts, faculties, talents, and predispositions, this still did not mean that he reduced these mental predispositions to physical organs in the brain. See also Ørsted op. cit. (6), p. 40. Ørsted op. cit. (3), vol. 2, p. 91. Henrich Steffens, Indledning til philosophiske Forelæsninger [1803], Johnny Kondrup, ed., Det Danske Sprog- og Litteraturselskab (Copenhagen: C. A. Reitzel, 1996). Harald Høffding, Danske Filosofer (Copenhagen: Gyldendal, 1909) pp. 29, 89. Hans Vammen, “Grundlaget for det Moderne Danmark. Hovedlinier i Dansk Politisk Idehistorie 1750–1850,” Historisk Tidsskrift, 84 (14. række bd. V), (1984), pp. 23–36. “Kritisk Romantik—Om Opfattelsen af den Danske Guldalder. I Anledning af en Disputats om N. L. Høyen,” Historisk Tidsskrift, 87 (15. række bd. II) (1987), pp. 18–38. Erik Schrøder, Frederik Christian Sibbern 1785– 1872. Politisk filosof og filosofisk politiker. Et studie i borgerlig bevidsthedsdannelse, Master’s Thesis (University of Copenhagen, 1977).

64

A. S. JACOBSEN

This is for example reflected in Sibbern’s philosophy. According to Sibbern, the spirit in nature also governed man and therefore it also ruled human faculties. Therefore, freedom was not an absolute concept since that would be in opposition to the demand of harmony in the individual as well as for the individual in relation to the whole, i.e. society or cosmos. The individual’s striving for freedom was subordinated the whole in such a way that he had free will on the one hand, but on the other it was subordinated natural order. In continuation hereof Sibbern stated the categorical imperative: you can do what you ought to do! In a political sense Sibbern employed this imperative to justify the conservation of the absolute monarchy. He believed that man had an obligation, both morally and ethically, to choose to become free. Everybody, despite actions that suggested the opposite, wished to be free, only conscience about freedom was unequally developed in different people.31 H. C. Ørsted’s views on will and morality echoed Sibbern’s. Ørsted also maintained the deep connection between nature and human mind at the same time regarding freedom as subordinated reason: Man is in an essential way an integrated link in the whole cosmos and subjected its laws, however with the great […] difference from all other earthly things that he can act with freedom within certain limits […] If we want a clear understanding of human life we must recall that we have freedom and that we are necessarily integrated parts of this whole. The apparent contradiction between these two truths is dissolved by the union of both into the more comprehensive truth that man is a member of an empire of reason.32

Moreover, the “spirit of freedom is aware of its inner connection with the legislation of reason, it is regulated accordingly without constraint.”33 Thus, like Sibbern, Ørsted believed in a virtuous way of life. Man had a free will to obey reason. Otherwise man would simply be subjected to his coarse and dirty senses with their depressing force opposite the rising force of the spirit. Whereas phrenologists found that will was subjected to motives and physical causes, Ørsted found that will was subordinated the whole i.e. the empire of divine reason which, however, was not considered to be a constraint. Different people had different talents and each individual should “aim at integrating oneself into the whole according to the virtues he is able to develop in himself.”34 While the economic and political crisis swept over Denmark during the first decades of the 19th century (the British fleet attacking Copenhagen in 1801; the violent British bombardment of Copenhagen in 1807; national bankruptcy in 1813, losing Norway in 1814, and a following agricultural crisis), focus of the upper classes shifted the country inward on the flourishing academic and literary culture. As a result the academic elite, the Romantics, acquired important positions at university, in the administration, and in the church and were involved in the influential academic societies and journals. The Romantics’ complete domi31

32 33 34

F. C. Sibbern, Om Forholdet imellem Sjæl og Legeme, saavel i Almindelighed som i phrenologisk, pathognomonisk, physiognomonisk og ethisk Henseende i Særdeleshed (Copenhagen: Schultz, 1849), p. 185. Schrøder op. cit. (30), p. 17. Ørsted op. cit. (6)., pp. 44–45. Ibid. p. 54. Ibid. pp. 57, 63.

PHRENOLOGY AND DANISH ROMANTICISM

65

nance of cultural life is revealed from the course of events of the many literary disputes of the period. In most cases the Romantics won the debates and the opponents were more or less dismissed from the respectable circles. There were internal conflicts of this power elite, but they did not appear openly in the press or in literature. Romanticism inherited from Germany constituted the majority opinion, and allied with the absolute monarchy and the limited press, the Romantics would attempt to suppress alternative ideas coming from French and particularly English literature in the 1820s.35 Among those who took Howitz’s side against the established Romantic minds in the freewill dispute was Carl Otto. He reviewed the literature in connection with the Howitz dispute in the French weekly periodical Messager Français du Nord, which was published by a Swedish Count Gyllembourg in Copenhagen in 1825. In pungent phrases Otto characterised the current situation in Denmark as a “literary slavery” because certain authorities completely dominated opinion and nobody dared to oppose their views. He characterised the dispute as “a beginning of the combat between enlightenment and obscurity, between facts and reveries, between reason and mysticism.”36 Otto’s own phrenological campaign led to extensive discussions of the subject in the journals and newspapers of the day such as Telegraphen (the telegraph), Kjøbenhavnsposten (the Copenhagen post), Messager Français du Nord, the leading literary journal Maanedsskrift for Litteratur (the monthly writings of literature), and in J. L. Heiberg’s satirical Kjøbenhavns flyvende Post (the Copenhagen flying post).37 Generally representatives of the academic elite were sceptical about the tenets of phrenology and the implications of it for the system of education, the philosophical question of free will, etc. They regretted Otto’s announcement of phrenology as a new empirically based philosophy of mind meant to replace prevailing ideas from German Naturphilosophie. In the journal Messager Français du Nord the fact that German philosophy was so dominant in Denmark was 35

36

37

Vammen 1984, 1987 op. cit. (30). In the period 1799–1848 the Danish absolute monarchy limited the freedom of the press. The result was a standstill of political debate until the introduction of the assembly of the Estates of the Realm [stænderforsamlingerne] in 1831 as a political institution opposed to the absolute monarchy. Until 1831 critical voices were rarely heard against the existing system as the control of the press was carefully enforced. The consequences of opposition were either exile, prison for life, or permanent censorship. Ove Hornby, “Through bankruptcy to democracy,” Danish Journal, special issue: Hans Christian Ørsted (Copenhagen: Schultz, 1977), pp. 6–13, on p. 11. Carl Otto, “La dispute en Danemarck sur la liberté morale,” Messager Français du Nord, 1 (1825), pp. 197–201, 212–213, on p. 198. The paper is anonymous, but in his memoirs Otto mentioned that he wrote it: Otto 1879 op. cit (21), pp. 186–87. F. G. Howitz, Ultimatum, angaaende Determinismus og Etatsraad Ørsteds Fortsatte Bemærkninger om samme (Copenhagen, 1825), pp. 50–51. Henrik Gerner von Schmidten, “Om Phrenologien i Anledning af Dr. Ottos Phrenologiske Tidsskrift,” Maanedsskrift for Litteratur, 1 (1828), pp. 101–110. J. L. Heiberg, “Organet for Saltsands,” Kjøbenhavns flyvende Post, No. 22., 1827. “Phrenologisk Opgave,” Kjøbenhavns flyvende Post, No. 32, 1828. J. C. Lange, “Om Phrænologien og adskillige dermed beslægtede Gjenstande, med nærmest Hensyn paa Dr. med. C. Otto’s derom udgivne Værk,” Telegraphen, 5(1), (1825), pp. 1–16, 78–80. Anonymous, “Spørgsmaale til Phrenologen,” Kjøbenhavnsposten, 1(70), (1827), pp. 282–283. “Dr. Galls Hjerneskal. Gaade for Phrenologerne,” Kjøbenhavnsposten,3(11), (1829), pp. 47–48. “Bemærkning i Anledning af den phrenologiske Gaade i No. 10 af Kjøbenhavnsposten,” Kjøbenhavnsposten, 3(13), (1829), p. 55. Carl Otto, “Svar fra Phrenologen,” Kjøbenhavnsposten, 1(71), (1827), p. 287.

66

A. S. JACOBSEN

explicitly nailed down as the reason that phrenology was more or less ignored here in comparison to in France.38 From the standpoint of the medical community the reaction against phrenology took place most actively at the Scandinavian Meetings for Natural Philosophers in 1844 and in 1847. At this time the dawning positivistic attitude among physicians based on the rapidly growing anatomical knowledge, led to the questioning of the scientific status of phrenology. However, a similar development in psychology, which led to this science’s distancing itself from philosophy and approaching more biological methods, led the professor of philosophy and psychology F. C. Sibbern to support Otto’s phrenological viewpoints at the same meeting. Sibbern was at this time working on a quite comprehensive treatise on soul and body, which contained a rather long chapter on phrenology.39 Given the previous sections it is surprising that Sibbern should be found among the “friends” of phrenology, since he was otherwise clearly an insider in the Copenhagen Romantic milieu. He was personally acquainted with Steffens and Schelling and his works were strongly influenced by Schelling’s philosophy of identity. Locally he was part of the inner circle around the Ørsted brothers.40 However, contrary to what might be expected from a Romantic, Sibbern endorsed faculty psychology, i.e. he believed in certain innate talents and predispositions both with regard to skills, feelings, and character. He found that Gall’s theory only corroborated the theory of faculty psychology and even willingly accepted the phrenologists’ view of the mind’s physical origin. In fact, Sibbern ended up rejecting that nature was vague and indeterministic, and stated his belief in a spirit and nature, which were material, mechanical, and deterministic in all respects. He stated that free mental movement [den frie aandelige Bevægelse] was based on a general mechanistic nature.41 This physical determinism was radically different from the view of the Naturphilosophen and the Romantics generally. In addition, Sibbern did not find, like the Naturphilosophen, that phrenology lacked a philosophical basis or that Gall was not philosophical in his approach. Nor did he find phrenology in opposition to the idea of unity of mind and nature even though it fragmented the brain and soul. On the contrary, Sibbern found that Gall’s doctrine “elucidated that the cause of all human thinking, all human aspiration and striving, is something internal a priori, or something acting in the original human nature.” Thus apparently he found perfect correspondence between phrenology and Naturphilosophie.42 38

39

40

41 42

Anonymous, “La Phrénologie ou la doctrine de Gall et de Spurzheim en Danemarck,” Messager Français du Nord, 1 (1825), pp. 136–140, 145–147, on p. 136 Forhandlinger ved de skandinaviske Naturforskeres fjerde Möde, i Christiania den 11–18 Juli 1844 (Christiania, 1847), pp. 51–63, 78–92. Forhandlinger ved de skandinaviske Naturforskeres femte Møde, der holdtes i Kiøbenhavn fra den 12te til den 17de Juli 1847 (Copenhagen: Gyldendal, 1849), pp. 178–204, 928–932. Sibbern op. cit. (31), pp. 171–419. For literature on Sibbern see Schrøder 1977 op. cit. (30). J. Himmelstrup, Sibbern. En monografi (Copenhagen: Schultz, 1934). Aase Nielsen, En sammenligning mellem ideindholdet i Steffens: Indledning til philosophiske Forelæsninger (1803), Sibbern: Om Erkjendelse og Granskning (1822) og Ørsted: Aanden i Naturen (1850), Master’s Thesis, Copenhagen University, 1957. Sibbern op. cit. (19), p. 48. Sibbern op. cit. (31), pp. 184–85, 414. Sibbern op. cit. (31), pp. 184, 198, 201–202, 261, 411. Hagner op. cit. (13). Trevor H. Levere, “S. T. Coleridge and the Human Sciences: Anthropology, Phrenology, and Mesmerism,” in Marsha P. Hanen, Margaret J. Osler, and Robert G. Weyant, eds., Science, Pseudo-science and Society (Waterloo, Ontario: Wilfried Laurier University Press, 1980) pp. 171–192, on pp. 184–85.

PHRENOLOGY AND DANISH ROMANTICISM

67

Sibbern distinguished between cranioscopy, organology, i.e. the proclamation of several separate cerebral organs and their localisation, and “this kind of psychology which […] has attached completely to the former doctrine of cerebral organs,” i.e. faculty psychology. As a psychologist, he was mainly interested in the latter, and did not give much attention to cranioscopy. He regretted the immense influence this part of phrenology had acquired and emphasised the need for new observations and experiments to support this discipline. Still he had no doubts whether skilled phrenologists were able to make the inferences correctly, and he even donated his own skull and brain for further research.43 Indeed, Sibbern’s cranium and brain were kept after his death and are still to be found at the Laboratory of Biological Anthropology at the Panum Institute, Copenhagen University (see photo of Sibbern’s cranium). Probably these remains were not subjected to a phrenological examination or Otto would surely have mentioned it. However, while still alive Sibbern should have let his son be investigated phrenologically by Otto. Otto felt quite honoured and proud that Sibbern, who was after all “a metaphysician,” nevertheless was in favour of phrenology. Otto suggested—and quite plausibly—that partly because Sibbern recognised phrenology, it had not aroused such a great opposition in Denmark, in comparison to what was the case elsewhere in Europe and in America.44 We must on the other hand presuppose that Sibbern’s quite long and dry treatise was read only by few contemporaries and therefore had a limited impact.

6. CONCLUSION Clearly the discrepancy between the philosophical principles and implications of phrenology and Romanticism outlined the picture of how Gall’s doctrine was received in Copenhagen and why phrenology did not gain ground here at its first appearance. The Romantics, i.e. the powerful intellectual elite in Copenhagen were philosophically biased against phrenology, while primarily people disagreeing with Romanticism were inclined to endorse phrenology. Thus, it was the philosophical perspectives of phrenology rather than the purely political, which were debated, and phrenology was never used as an instrument in reform discussions as seems to have been the case in Edinburgh for instance.45 The Naturphilosophen regarded phrenology the same way as they regarded empirical science in general: it could teach us nothing about the spirit in nature and in man. Generally Romantics would hold that soul and mind could not be fragmented and reduced to specific, localised physical organs, and soul, mind, the brain and the body acted as a whole. However, naturally there are exceptions to this generalisation of the romantics versus the phrenologists. Not all Romantics

43 44 45

Sibbern op. cit. (31), pp. 182–83. Otto op. cit. (15), pp. 368–369. Otto 1879 op. cit. (21), p. 208. Otto to Sibbern 1849 op. cit (24). Steven Shapin, “Phrenological Knowledge and the Social Structure of early Nineteenth-Century Edinburgh,” Annals of Science, 32 (1975), pp. 219–243.

68

A. S. JACOBSEN

would deny the presence of individual mental faculties whether innate or subjected to education. Thus the otherwise Romantic professor of philosophy F. C. Sibbern willingly accepted such an idea. Despite the fact that Sibbern was clearly inspired by Schelling’s philosophy of identity, he even accepted the existence of material conditions in terms of physical organs behind the mental faculties. Thus, there did not exist incommensurability between Romantics and phrenologists.46 Roskilde University

ACKNOWLEDGMENTS I would like to thank Pia Bennike at the Anthropological Laboratorium, Copenhagen University, Frank Allan Rasmussen, Ion Meyer, and Rikke Claësson at the Medical Museion, Copenhagen, and Svend Larsen at the State and University Library, Århus, for kind assistance with images. I further thank the History of Danish Science project, at the History of Science Department, University of Aarhus, for supporting this research.

46

Geoffrey N. Cantor, “The Edinburgh Phrenology Debate: 1803–1828,” Annals of Science (1975), 32, pp. 195–218.

PHRENOLOGY AND DANISH ROMANTICISM

The young Hans Christian Ørsted (1777–1851). (Courtesy the Royal Library, Copenhagen)

69

70

A. S. JACOBSEN

Phrenological head. (Courtesy the Medical Museion, Copenhagen)

PHRENOLOGY AND DANISH ROMANTICISM

Carl Otto (1795–1879). (Courtesy the State and University Library, Aarhus)

71

72

A. S. JACOBSEN

Phrenological skull most likely originating from Carl Otto’s collection. (Courtesy the Laboratory of Biological Anthropology, the Panum Institute, Copenhagen University)

PHRENOLOGY AND DANISH ROMANTICISM

Frederik Christian Sibbern (1785–1872). (Courtesy the State and University Library, Aarhus)

73

74

A. S. JACOBSEN

F. C. Sibbern’s cranium. (Courtesy the Laboratory of Biological Anthropology, the Panum Institute, Copenhagen University)

NATURAL ENDS AND THE END OF NATURE Naturalizing Kant’s teleology PAUL GUYER

1. SALVAGING TELEOLOGY Kant’s Critique of the Teleological Power of Judgment was clearly his response to Hume’s Dialogues Concerning Natural Religion. Hume’s Dialogues seem to be a withering criticism of the argument from design, yet Kant apparently wanted to salvage something from the wreckage of traditional teleology. It may be hard to see why he would have thought he could do so until we see that Hume’s attack is not in fact a complete rejection of traditional teleology. Once we see this, we can understand both how Kant could have thought that he could salvage something from traditional teleology, but also why he would have thought that Hume’s critique was incomplete. What Hume actually argued is that it is natural and unavoidable for human beings to think of the world as the intentional product of an intelligent author, but that all speculation into the character and purposes of this author is completely groundless. What Kant does in his Critique is to accept Hume’s supposition that it is natural for us to posit an intelligent author of nature, accept Hume’s claim that we have no basis for a theoretical determination of the character and purposes of such an author, but add that it is nevertheless incoherent for us to conceive of an intelligent yet purposeless author of nature. We cannot conceive of nature as having a design without also conceiving of it as having a point. To fill this gap, he holds that the only end that we must represent as of unconditional value, namely the full development of our own morality, is also the only thing that we can conceive of as the final end of the creation of nature and thus the ultimate end of nature itself. In a word, he accepts Hume’s rejection of an anthropomorphic conception of God, yet argues that we must replace that with an anthropocentric, although morally anthropocentric, conception of nature. Since this conception does not amount to theoretical cognition, yet unlike a postulate of pure practical reason it characterizes not just the author of nature but nature itself, the same thing that is the object of our theoretical cognition, Kant attributes this conception to the power of judgment, but to the reflecting rather than the determining power of judgment.

75 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 75–96. © 2007 Springer.

76

P. GUYER

In this paper, I briefly characterize the chain of thought by which Kant argues that it is natural for us to conceive of the world in this morally anthropocentric way, and then ask what if anything in Kant’s teleological outlook could survive the drastic change that our conception of nature has undergone in the two centuries since Kant wrote. We can begin by reminding ourselves of the character of the traditional teleology to which both Hume and Kant were responding. One example of traditional teleology that Kant, although not Hume, would have known is Christian Wolff’s Rational Thoughts on the Intentions of Natural Things, first published in 1723. Wolff was one of the great proponents of the early Enlightenment in Germany. Yet Wolff confidently argued that we know two definite things about the purpose of the world and everything in it: first, that the world was created by God in order to reveal his greatness to us, and, second, that beyond teaching us this lesson, everything else in the world was created was created for our own use and happiness. Thus, Wolff first asserted that [t]he chief aim of the world is that we should know God’s perfection from it. Now if this is what God wanted to achieve, then he had to arrange the world in such a way that a rational being could draw from consideration of these grounds his attributes and infer with certainty everything else that one can know of him.1

Wolff then proceeded to explain how various global characteristics of the created world are a mirror in which we can come to see the perfections of God. Thus, for example, the sheer number of things in the world is “a mirror of the infinite cognition of God”;2 “the connection of the things in the world to each other” by the most direct possible routes is a “mirror of his wisdom”;3 and even the contingency of the existence of the particular things in the world is a mirror of God’s freedom: “If the world were necessary, then we could no longer know from it that there is a God, that is, a being distinct from it in which the ground of its reality is to be found,”4 for it is precisely “the contingency of the world” that “makes it into a mirror of the freedom of the divine will.”5 Wolff expresses the second main thesis of his teleology, that the world is created for the purposes of man, in language like this: “One cannot say otherwise than that God made the earth so that it would be occupied, and on that account arranged everything in it so that it would be fit as a dwelling for men and animals. Man finds here everything that may be suitable to him for his nourishment, dress, dwelling, attainment of science and art, and whatever is necessary for the fulfillment of his duties.”6 Human happiness is not only man’s end, as Wolff had argued in his moral philosophy, but also God’s end for man. Both Hume and Kant rejected such confident speculation about God’s intentions in the creation of the world. In the Dialogues, Hume’s spokesman Philo, with unsurpassed British understatement, objects to similar claims put in the mouth of the character Cleanthes: 1

2 3 4 5 6

Christian Wolff, Vernünfftige Gedancken von der Absichten der natürlichen Dinge, 2nd ed. (Leipzig and Frankfurt: Renger, 1726), ch. II, §8, p. 6. Ibid. ch. II, §13, p. 16. Ibid. ch. II, §14, p. 19. Ibid. ch. II, §9, p. 7. Ibid. ch. II, §11, p. 12. Ibid. ch. VII, §66, p. 97.

NATURAL ENDS AND THE END OF NATURE

77

But when we look beyond human Affairs and the Properties of the surrounding Bodies: When7 we carry our speculations into the two Eternities, before and after the present State of things; into the Creation and Formation of the Universe; the Existence and Properties of Spirits; the Powers and Operations of one universal Spirit, existing without Beginning and without End; omnipotent, omniscient, immutable, infinite, and incomprehensible: We must be far removed from the smallest Tendency to Scepticism not to be apprehensive, that we here got quite beyond the Reach of our Faculties.8

It is easy to read this as a complete rejection of any teleological speculation. However, that seems to leave Philo’s apparent concession to an argument from design in the final part of the Dialogues as a cowardly retraction of all that has gone before: A Purpose, an Intention, a Design strikes every where the most careless, the most stupid Thinker; and no man can be so harden’d in absurd Systems, as at all times to reject it. That Nature does nothing in vain, is a Maxim establish’d in all the Schools, merely from the Contemplation of the Works of Nature, without any religious Purpose; and, from a firm Conviction of its Truth, an Anatomist, who had observ’d a new Organ or Canal, wou’d never be satisfy’d, till he had also discover’d its Use and Intention. One great Foundation of the Copernican System is the Maxim, that Nature acts by the simplest Methods, and chooses the most proper Means to any End; and Astronomers often, without thinking of it, lay this strong Foundation of Piety and Religion. The same thing is observable in other Parts of Philosophy: And thus all the Sciences almost lead us insensibly to acknowledge a first intelligent Author; and their Authority is often so much the greater, as they do not directly profess that Intention.9

It is hard to believe that Kant did not have this passage in mind when he wrote himself of the indispensability of the maxims of the reflecting power of judgment in the third Critique.10 Before we can turn to Kant, however, we must ask how Philo, and presumably Hume speaking through him, could have delivered both of the passages that have just been quoted. An explanation has recently been offered by H.O. Mounce. On his account, Philo’s argument with Cleanthes in the first part of the Dialogues (sections 2 through 8) is a rejection of “the argument from design [as] an ordinary empirical argument which, in establishing the existence of God, also reveals his nature as analogous to our own, differing only in degree”—“The God whom Cleanthes affirms is anthropomorphic… the God, Hume believes, of ordinary religion.”11 Cleanthes’s affirmation of an anthropomorphic God is evident in passages such as this: I shall briefly explain how I conceive this Matter. Look round the World: Contemplate the Whole and every Part of it: You will find it to be nothing but one great Machine, subdivided into an infinite Number of lesser Machines…All these various Machines, and even their most minute Parts, are adjusted to each other with an Accuracy, which ravishes into Admiration all Men, who have ever contemplated them. The curious adapting of Means to Ends, throughout all Nature, resembles exactly, tho’ it much exceeds, the Productions of human Contrivances; of human Design, Thought,

7 8

9 10 11

Hume, Dialogues, [?] Dialogues concerning Natural Religion, Part I; in David Hume, The Natural History of Religions and Dialogues concerning Natural Religion, edited by A. Wayne Colver and John Valdimir Price (Oxford: Clarendon Press, 1976), pp. 151–152. Dialogues, Colver and Price, p. 245. For example, CPJ, Introduction, section IV, 20: 210–11. H. O. Mounce, Hume’s Naturalism (London and New York: Routledge, 1999), p. 116.

78

P. GUYER Wisdom, and Intelligence. Since therefore the Effects resemble each other, we are led to infer, by all the Rules of Analogy, that the Causes also resemble; and that the Author of Nature is somewhat similar to the Mind of Man; tho’ possessed of much larger Faculties, proportion’d to the Grandeur of the Work, which he has executed.12

This is the sort of claim that Philo makes into mincemeat, arguing that if we are to take such an analogy seriously then perhaps we should conclude that the world as we observe it was actually made by an immature or a superannuated god, or a committee of gods, or that it grew like a vegetable rather than being built like a machine, and so on. Nevertheless, as Mounce observes, Philo, like his own intelligent author, “is not advocating extreme scepticism… What he advocates is mitigated not extreme scepticism.”13 What this means is that Hume in fact allows that it is natural for us to assume that the universe as a whole has some cause, and indeed some cause proportionate to its magnitude and thus adequate to produce it, although since we have not observed that cause, let alone observed it repeatedly, it “transcends [our] experience”14 and we have no basis for a determinate conception of it. What Mounce does not spell out, but needs to make this interpretation work, is the lemma that it is natural and permissible for us to believe that every object or event has some cause, even when we have not been or cannot have been in a position to observe the succession of that specific kind of event upon its cause repeatedly or even once. However, Hume clearly does believe that he can explain why “we pronounce it necessary, that everything whose existence has a beginning, shou’d also have a cause,” even though, as he says, he will find it “more convenient to sink” his answer to “this question” into his answer to the question “Why we conclude, that such particular causes must necessarily have such particular effects.”15 Once having sunk the question of our belief in the general principle that every event has some cause into that of our belief in particular causal connections, Hume never does bring it back to the surface of his argument again; but presumably his thought is that once we have had repeated experience of various particular causal connections, it is natural for us to abstract from their particularity and form a general expectation of some suitable contiguous and antecedent cause every time we experience a new sort of object or event (although, to be sure, this might run afoul of his Berkeleian worries about abstract ideas). Once armed with this lemma, Hume could then indeed have concluded, as Philo does, that some sort of belief in an author of the works of nature is both natural, and, as long as no baseless claims about its nature are made, permissible: So little, reply’d Philo, do I esteem this Suspence of Judgment in the present Case to be possible, that I am apt to suspect there enters somewhat of a Dispute of Words into this Controversy, more than is usually imagin’d. That the Works of Nature bear a great Analogy to the Products of Art is evident; and, according to all the Rules of good Reasoning, we ought to infer, if we argue at all concerning them, that their Causes have a proportional Analogy. But as there are also considerable Differences,

12 13 14 15

Dialogues, Colver and Price, pp. 161–162. Mounce, p. 117. Mounce, p. 116. Hume, A Treatise of Human Nature, Book I, Part III, sections ii and iii; in the edition by L. A. SelbyBigge, rev. P. H. Nidditch (Oxford: Clarendon Press, 1975), pp. 78, 82.

NATURAL ENDS AND THE END OF NATURE

79

we have reason to suppose a proportional Difference in the Causes; and in particular ought to attribute a much higher Degree of Power and Energy to the supreme Cause than any we have ever observ’d in Mankind. Here then the Existence of a DEITY is plainly ascertain’d by reason.16

On Hume’s account, belief in both particular causal connections and the general principle that every event has some cause is not due to an operation of the understanding, confined as that faculty is to the logic of identity and noncontradiction, but is nevertheless entirely natural. The principle that a cause must always be proportionate to its effect is also natural—“When we infer any particular cause from an effect, we must proportion the one to the other.”17 And thus the inference to an indeterminate author of the universe, although not to an anthropomorphic God, is entirely natural and indeed unavoidable. In the Critique of Pure Reason, of course, Kant argues that our belief in the general principle of causality is more than merely natural and unavoidable. We can see him as arguing in the Critique of Teleological Judgment, however, that belief in a special type of causation, namely an intelligent creation of organisms and in analogy to them all of nature, is natural and unavoidable in Hume’s sense, even if not necessary in the sense of the first Critique. However, we can also see him as going on to claim that Hume’s argument is incomplete. On Kant’s account, if it is natural and unavoidable for us to conceive of nature or any of its parts as the product of intelligent design, then it is also natural and unavoidable for us to conceive of that creation as purposive, that is, as having a point. He then further agrees with Hume that we have no ground for determinate theoretical speculation about the character and thus the purposes of such an intelligent and purposive author of nature. However, that does not leave us without resources for conceiving of the point of nature: we have a conception of a final end of unconditional value available to us, namely the full development of our own morality, and we naturally can and do use this conception in order to conceive of a point for nature. Kant thus accepts Hume’s rejection of theoretical anthropomorphism, but replaces the gap left by that with his own moral anthropocentrism. To understand Kant’s reconstruction of teleology, we should see it as proceeding in three stages. First, the teleological perspective must be shown to be natural. Second, it must be shown to be free of contradiction. Finally, it must be shown to be beneficial, to have a regulative use that is genuinely useful in the prosecution of inquiry, in the execution of action, or in both. These are the stages through which Kant’s defense of pure reason proceeds in the Transcendental Dialectic of the first Critique, and we can understand his argument as proceeding through similar stages in the Critique of Teleological Judgment. In the first Critique, the conceptions at issue are, of course, the three Ideas of pure reason, namely the idea of the unitary self, the idea of the world-whole, and the idea of God. The first stage of Kant’s argument, in the First Book of the Transcendental Dialectic, is to show how these

16 17

Dialogues, Part 12, Colver and Price, pp. 247–248. Enquiry concerning Human Understanding, Section 11; in David Hume, An Enquiry concerning Human Understanding: A Critical Edition, edited by Tom L. Beauchamp (Oxford: Clarendon Press, 2000), p. 102.

80

P. GUYER

concepts of pure reason are naturally generated from the relational categories of the understanding by the iteration of syllogistic inference. In the Second Book of the Dialectic, Kant argues that these Ideas lead to fallacies and contradictions if they are understood as a constitutive source for theoretical cognition of things as they are in themselves. But in the Appendix to the Transcendental Dialectic, Kant argues, as the second positive stage of his overall argument, that these Ideas can be understood without fallacy or self-contradiction (antinomy) if they are interpreted regulatively rather than constitutively, and then, as the third stage of his argument, that these Ideas have a beneficial use. Kant explicitly associates these latter two stages of his argument with the idea of a “transcendental deduction” of the ideas of pure reason, although not a “deduction of the same kind as the categories” (A 669–670/B 697–698), arguing that for a proper deduction of these ideas it must first be shown that “there is not the least thing to hinder us from assuming these ideas,” and that if there appears to be something that would hinder us from assuming them, such as antinomies, these must be resolved before we can proceed (A 673/B 701), and then that it must also be shown that they have a beneficial use, for instance “as regulative principles for the systematic unity of the manifold of empirical cognition in general, through which this cognition, within its proper boundaries, is cultivated and corrected more than could happen without such ideas, through the mere use of the principles of understanding” (A 671/B 699); but both of these steps presuppose that the ideas in question have first been shown to be natural products of human reason. The conditions for a deduction of the regulative use of the ideas of pure reason are summed up thus: Now if one can show that although the three kinds of transcendental ideas…cannot be referred directly to any object corresponding to them and to its determination, and nevertheless that all rules of the empirical use of reason under the presupposition of such an object in the idea lead to systematic unity, always extending the cognition of experience but never going contrary to experience, then it is a necessary maxim of reason to proceed in accordance with such ideas. (A 671/B 699)

Kant parallels these latter two steps in a deduction of Ideas when he explains what he has in mind by a transcendental deduction of the regulative principles of reflecting judgment in Section V of the Introduction to the third Critique. Here too he emphasizes that in order to provide a deduction of a regulative principle we must first provide a nonproblematic interpretation of our concept, in this case the concept of purposiveness, and then demonstrate its utility. Thus, he argues first that “this transcendental concept of a purposiveness of nature is neither a concept of nature nor a concept of freedom, since it attributes nothing at all to the object (of nature), but rather only represents the unique way in which we must proceed in reflection on the objects if nature with the aim of a thoroughly interconnected experience” (CPJ, 5: 184), and then argues that this concept is useful, “because without presupposing it, we would have no order of nature in accordance with empirical laws, hence no guidelines for an experience of this in all its multiplicity and for research into it” (5: 185). We can make sense of the structure of Kant’s overall argument in the Critique of Teleological Judgment in light of this analysis. First, Kant shows why it is natural to introduce the concept of purposiveness into our thought about nature.

NATURAL ENDS AND THE END OF NATURE

81

This stage of his argument begins by insisting upon the special character of our experience of organisms, but then goes on to claim that once we have been led to conceive of organisms as products of purposive design it is only natural for us to conceive of the whole of nature in that way too. Next, Kant shows how we can conceive of both organisms within nature and nature as a whole in this way without contradiction; not surprisingly, he includes an “antinomy” of teleological judgment and its resolution at this stage of his argument. Finally, Kant shows what the use of a teleological conception of nature is. That the conception of nature as purposive has a heuristic role in the pursuit of empirical knowledge of nature is in fact emphasized from the outset of Kant’s Introduction to the whole Critique of the Power of Judgment; but the “methodology” of the teleological power of judgment adds to that an account of the moral significance of the conception of nature as a purposive system. This can be thought of as the final stage of Kant’s justification or “deduction” of his revision of traditional teleology. 2. NATURAL ENDS AND THE END OF NATURE I have discussed the details of Kant’s arguments elsewhere, and in this paper I can merely outline them.18 Kant’s philosophy as a whole is thoroughly teleological in the sense that he always tries to show, and assumes that he can show, that there is nothing in nature, thus nothing in our own nature, that is simply idle and useless, as long as it is properly understood. He thus wants to argue that the idea that nature is purposive, an idea to which, he takes it, we are naturally prone, must also have a proper use. The idea that nature is purposive is the idea that it has the kind of organization that could have been given to it only by an intelligent author, through a concept of its form as antecedent efficient cause of its actual form, and that since such a author must be conceived of as not merely intelligent but also rational, nature must have not only an intentional form but also an aim or a point, in Kant’s terminology a purpose or an end (Zweck). In this passage from the Introduction to the third Critique Kant both defines his concept of purposiveness and claims that the principle of purposiveness is the fundamental principle of the power of (reflecting) judgment: Now since the concept of an object insofar as it at the same time contains the ground of the reality of this object is called an end, and the correspondence of a thing with that constitution of things that is possible only in accordance with ends is called the purposiveness of its form, thus the principle of the power of judgment in regard to the form of things in nature under empirical laws in general is the purposiveness of nature in its multiplicity. (CPJ, Introduction IV, 5:180)

18

See especially “The Unity of Nature and Freedom: Kant’s Conception of the System of Philosophy,” in Sally Sedgwick, ed., The Reception of Kant’s Critical Philosophy (Cambridge: Cambridge University Press, 2000), pp. 19–53;“Organisms and the Unity of Science,” in Eric Watkins, ed., Kant and the Sciences (Oxford: Oxford University Press, 2001), pp. 259–281; and“From Nature to Morality: Kant’s New Argument in the ‘Critique of Teleological Judgment’,” in Hans Friedrich Fulda and Jürgen Stolzenberg, eds., Architektonik und System in der Philosophie Kants (Hamburg: Felix Meiner Verlag, 2001), pp. 375–404.

82

P. GUYER

He insists that the idea of the purposiveness of nature is not a source of theoretical insight, but that it nevertheless has a valuable and indeed indispensable heuristic role for us: The purposiveness of nature is thus a special a priori concept that has its origin strictly in the reflecting power of judgment. For we cannot ascribe to the products of nature anything like a relation of nature in them to ends, but can only use this concept in order to reflect on the connection of appearances in nature that are given in accordance with empirical laws. (Ibid. 5: 181)

In fact, Kant argues that the concept of the purposiveness of nature has several indispensable heuristic roles. First, he argues that the idea that our concepts of the kinds of things there are in nature and the natural laws that govern their behavior constitute a system, and thus the idea that the things in nature themselves constitute a system, is vital for discovering the particular concepts of the kinds of things in nature and their laws, as well as for comprehending the lawlikeness, that is, the necessity, of those particular laws.19 Second, he argues that the idea of particular things in nature, namely, organisms, as systems in which each part of the thing is both cause and effect of the whole, is indispensable even for our search for mechanical explanations of the functions of such things, although it may also mark the limits of the possibility of our success in searching for such explanations. And finally, Kant argues that once again the idea of the things of nature as a whole as comprising a single system, although in this case an idea of the systematicity of nature as a whole to which we are led by our idea of the systematicity of organisms as particular things in nature, has an indispensable role not in our conduct of science but rather in morality. Kant argues that if we are to be able to conceive of nature as a whole as a single system, we must be able to conceive of some determinate and unique point to it, but that such a point, or as he calls it “ultimate end,” can only be something of unconditional value, or as he calls it a “final end.” But the only thing that has that sort of value is human morality itself. So if we are to conceive of nature as a whole as a single system, we can only do so by conceiving of the full development of human morality—certainly not human happiness, or at least not human happiness alone and for its own sake—as its point. This has a value in driving home to us the importance of morality, even when we start from a point in the conduct of natural science that has nothing to do with morality. It has the point of confirming us in our assumption that nature must be a fit arena for the development of human morality, an assumption that we are also forced to reach from within morality itself, through the concept of the highest good. But it also has the point of reminding us that not merely must nature be compatible with morality, but also that morality must be compatible with nature: that is, that even those goals set by morality itself must be able to be realized in a way that respects the integrity of nature as a system.

19

I have explored these thoughts more fully in “Reason and Reflective Judgment: Kant on the Significance of Systematicity,” Nous 24 (1990): 17–43;“Kant’s Conception of Empirical Law,” Proceedings of the Aristotelian Society, Supplementary Volume 64 (London: The Aristotelian Society, 1990), pp. 221–242; and, most recently, “Two Puzzles about Kant on the Systematicity of Nature” (not yet published).

NATURAL ENDS AND THE END OF NATURE

83

Kant’s argument about teleology thus consists of three main stages. At the first stage, Kant appeals to two different grounds to convince us that we must conceive of nature as purposive, although this conception is compelling only for reflecting judgment. The second stage of the argument is to show that this conception of nature is coherent. But to conceive of nature as purposive means not just that we conceive of it as if it were created in accordance with a plan, but also as if it has a point; and at the third stage of his argument, Kant thus argues that we must conceive of the point of nature as not only our own cognitive satisfaction but also the full development of our morality. 2.1. First stage In the first stage of his overall argument, Kant appeals to two different sorts of consideration, general considerations about the systematicity of our concepts of nature on the one hand and considerations about our specific experience of organisms as things in nature on the other. (a) The introduction to the book as a whole focuses on the first issue. Here Kant argues that the “universal transcendental concepts of nature” and the “universal transcendental laws” that flow from them, that is, the categories of understanding and the principles of empirical thought that are derived from them in the Critique of Pure Reason, are not sufficient to determine a unique and necessary set of particular laws of nature, presumably because alternative sets of particular laws of nature could be equally consistent with both our actual empirical intuitions of the objects of nature and the general constraints of the pure categories of the understanding and the principles of judgment. Yet, Kant holds, in order for us to have knowledge of the particular laws of nature, these laws must be not only determinate but at least in some sense necessary. He claims: There is such a manifold of forms in nature, as it were so many modifications of the universal transcendental concepts of nature that are left undetermined by those laws that the pure understanding gives a priori, since these pertain only to the possibility of a nature (as object of the senses) in general, that there must nevertheless also be laws for it which, as empirical, may seem to be contingent in accordance with the insight of our understanding, but which, if they are to be called laws (as is also required by the concept of a nature), must be regarded as necessary on a principle of the unity of the manifold, even if that principle is unknown to us. (CPJ, Introduction, IV, 5: 179–180)

He then suggests that the necessity of particular laws of nature can be understood if these laws are regarded as part of a unique system of laws, but that we can only conceive of such a system as if it were the product of an intelligence, greater than our own, which has produced nature in accordance with a plan which has as its aim (at least) the satisfaction of our own cognitive objectives: Now this principle can be nothing other than this: that since universal laws of nature have their ground in our understanding, which prescribes them to nature (although only in accordance with the universal concept of it as nature), the particular empirical laws, in regard to that which is left undetermined in them by the former, must be considered in terms of the sort of unity they would have if an understanding (even if not ours) had likewise given them for the sake of our faculty of cognition, in order to make possible a system of experience in accordance with particular laws of nature.

84

P. GUYER

This argument depends not only on the assumption that the categories alone underdetermine our choice of particular empirical concepts and laws of nature, but also on the assumption that the categories alone are not sufficient to ground the necessity of particular laws of nature. For Kant’s general principle is not that any form of order must be the product of thought, but rather that anything that we can know to be universal and necessary, hence that we must know a priori, must be the product of our own thought. It is thus because we assume that the particular laws of nature are in some sense necessary, not just determinate, that we must conceive of them as if they were the product of intelligence, although presumably their necessity also implies their determinacy. Of course, since the intelligence posited in this case is not our own, we do not actually get a priori knowledge of particular laws of nature, and thus we get something more like the appearance of the necessity of such laws than genuine necessity. But apparently Kant thinks this is better than nothing. (b) Kant’s second consideration at this first stage of his overall argument is based on allegedly fundamental distinctions between organisms and the inorganic in nature. Kant begins the Critique of Teleological Judgment by explaining why we must conceive of organisms as “physical” or “natural ends.” He argues first that there is nothing in our experience of nature that on the face of it forces us to attribute “external” or “relative purposiveness” to things, that is, to suppose that they have been created because they have some specific end, such as being useful to creatures like ourselves. It may seem as if large sea-mammals and driftwood have been created for the benefit of humans living in arctic regions, but as soon as one asks “why human beings have to live” in those otherwise barren regions at all, any easy assumption of purposiveness here collapses (CPJ, §63, 5: 369). However, the experience of those specific things in nature that we call organisms forces the thought of purposiveness upon us, though in this case the thought of “internal” rather than “external purposiveness.” This is because we have to conceive of an organism as something that is both “cause and effect of itself,” which given the limitations of our usual way of thinking about causation, in which something cannot be both cause and effect of itself, is possible only if we think of an organism as a “natural end,” that is, something like a product of an antecedent conception of itself in nature, something like, although of course not exactly like, a product of human art (CPJ, §64, 5: 370)—an object is an end, remember, insofar as the ground of its reality is considered to lie in a concept of it (CPJ, Introduction IV, 5: 180). Specifically, Kant argues that there are a variety of processes typical of organisms in which the organism is considered to be both cause and effect of itself. He gives three examples: reproduction, in which one organism, by producing more of its kind, “unceasingly produces itself…and continuously preserves itself, as species”; growth, in which an organism as a whole takes in materials that it then prepares to become parts of itself; and finally self-preservation and maintenance, in which the whole depends upon parts that in turn depend upon the whole, as for example the life of a tree depends upon the functions of its leaves, which however depend upon the tree as a whole for their own existence (5: 371–372). The basis for Kant’s claim that these sorts of processes defeat our ordinary conception of causation is the assumption that in ordinary efficient causation, or what Kant calls

NATURAL ENDS AND THE END OF NATURE

85

mechanical causation, we always understand the character of a whole as the product of the antecedent character of its parts, and not vice versa (CPJ, §65, 5: 372). The only way that we can understand a whole to be the cause of its parts is then if we conceive of an antecedent concept of the whole as the cause of the parts that then come to constitute the whole. We understand the products of intentional human design and production in this way: for example, we understand a finished house as the product of an antecedent plan for the house which has led to the manufacture and collection of the parts which are then assembled into the finished product. In such a case, of course, we assume that there is not only an antecedent plan for the house, but also a purpose for the construction of house, such as renting it out for income (5: 372). That is, in the case of human productive activity we assume that efficient causation is a product of both a formal and a final cause. In order to have any sort of explanation of organisms, we then think of them as if they were the products of something like intentional production too, although one that is not our own and is not exactly like our own artistic production—for organisms have powers of self-repair and reproduction that our products, such as watches, do not (5: 374). But our assumption that where there is a formal cause there must also be a final cause is carried over to the case of organisms—that is what will be the basis for the subsequent stage of Kant’s argument, that we can only conceive of nature as purposive if we conceive it to have a point. Kant next suggests that the conception of purposiveness that is forced upon us by our experience of organisms naturally leads us to look at the whole of nature— not just the collection of our concepts of nature, as in the introduction—as a purposive system: It is therefore only matter insofar as it is organized that necessarily carries with the concept of itself as a natural end, since its specific form is at the same time a product of nature. However, this concept necessarily leads to the idea of the whole of nature as a system in accordance with the rule of ends, to which idea all of the mechanism of nature in accordance with principles of reason must now be subordinated (at least in order to test natural appearances by this idea). (CPJ, §67, 5: 378–379)

This passage suggests two different accounts of the generalization of the idea of purposiveness to all of nature. On the one hand, Kant’s concluding remark suggests that we simply try out the idea of the whole of nature as a single system, as a heuristic principle, in order to see where it leads, as does his statement in the next paragraph that “by its means we acquire only a guideline for considering things in nature, in relation to a determining ground that is already given, in accordance with a new, lawful order, and for extending natural science in accordance with another principle, namely that of final causes, yet without harm to the mechanism of nature” (5: 379). Perhaps the idea is that we simply have a natural tendency to explore the heuristic potential of this viewpoint. However, Kant’s opening statement that the concept of particular organisms as natural ends “necessarily leads to the idea of the whole of nature as a system in accordance with the rule of natural ends” suggests a stronger basis for the generalization, as does the concluding paragraph of the present section: In this section we have meant to say nothing except that once we have discovered in nature a capacity for bringing forth products that can only be conceived by us in

86

P. GUYER accordance with the concept of final causes, we may go further and also judge to belong to a system of ends even those things (or their relation, however purposive) which do not make it necessary to seek another principle of their possibility beyond the mechanism of blindly acting causes; because the former idea already, as far as its ground is concerned, leads us beyond the sensible world, and the unity of the supersensible principle must then be considered as valid in the same way not merely for certain species of natural beings but for the whole of nature as a system. (5: 380–381)

This suggests that if we are to posit a supersensible ground for organisms we must posit a single supersensible ground for everything in nature, because the idea of a supersensible ground is an idea of reason and such an idea necessarily connotes unity. (Hume’s Philo might have objected precisely to this assumption that the cause of nature as a whole must be unitary, since the causes we experience within nature certainly are not. Presumably Kant would have replied that the idea of a single supersensible ground of nature is a natural idea of human reason which leads to no illusion as long as it is understood regulatively rather than constitutively.) 2.2. Second stage In the second stage of his overall argument, which we can locate in the antinomy of teleological judgment, Kant aims to show how the idea of an intelligent and purposive author of nature can be coherently conceived. Kant begins this section by insisting on the distinction between mechanical explanation and explanation by final causes that seems necessary in the case of organisms, but ends up by demonstrating that both kinds of causation can be applied to all phenomena in nature, as long as they are understood as working at different levels. This argument can also be understood as a defense of our natural transition from the idea of organisms as purposive to the idea of nature as a whole as a purposive system, for it will show that the only way in which organisms can coherently be conceived as purposive necessarily leaves room for a conception of the whole of nature as purposive. Kant begins the antinomy by contrasting two “maxims” for the power of judgment that would appear to be incompatible: the “thesis” that “All generation of material things and their forms must be judged as possible in accordance with merely mechanical laws” and the “antithesis” that “Some products of material nature cannot be judged as possible according to merely mechanical laws,” rather, “judging them requires an entirely different law of causality, namely that of final causes” (CPJ, §70, 5: 387). Kant argues that if these were understood “as objective principles for the determining power of judgment, they would contradict one another” (5: 387), but if they are both understood merely as regulative principles for the reflecting power of judgment, then we could interpret the thesis to tell us to “pursue” mechanical explanation “as far as one can,” while the antithesis reminds us that nevertheless there will still be something, at least in the case of organisms, that will sooner or later resist our efforts at mechanical explanation (5: 388). But this is only a claim about the limits of “human reason,” not about the nature of things, and thus there is no actual contradiction between the two maxims. In the next section, however, Kant states that this approach is only a “preparation for the resolution” of the antinomy (CPJ, §71, 5: 388), and he then offers a more elaborate resolution. This is needed at least in part because treating

NATURAL ENDS AND THE END OF NATURE

87

the imperative to look for mechanical explanations of things in nature as a mere maxim can lead us to appeal to final causes too soon, and in the wrong way, as actually present within nature: “It must therefore be a certain presentiment of our reason, or a hint as it were given to us by nature, that we could by means of that concept of final causes step beyond nature and even connect it to the highest point in the series of causes if we were to abandon research into nature (even though we have not gotten very far in that)” (CPJ, §72, 5: 390). So what Kant now argues is that the antinomy of teleological judgment, like his other antinomies, can only be solved by appeal to the distinction between the sensible and the supersensible: the only proper way to conceive of a principle of final causes in the case of organisms is to conceive of it as a supersensible cause lying beneath or behind nature, not as part of the series of causes and effects within nature itself. But once we have conceived of the principle of final causes in that way, there is actually no further need for us to posit an insuperable distinction between organisms and the inorganic, even if we still believe, for whatever reason, that there are certain aspects of organisms that will in fact resist our efforts at mechanical explanation. Experience of a distinction between organisms and the inorganic may lead us to introduce the idea of final causes, but will not in fact be necessary to sustain it. Kant argues himself to this conclusion by contrasting the “idealism” and the “realism” of purposiveness in natural ends, where the idealism of such purposiveness ends up reducing it to a mere appearance and realism does not. He then introduces subsidiary forms of each of these main approaches to final causes. The idealism of purposiveness can take the form either of its “casuality” or of its “fatality.” The doctrine of casuality, which Kant associates with Epicurus, is the idea that all organization of things, including any apparently purposive organization, is the product of mere chance, random collisions among the atoms. The doctrine of fatality, which Kant associates with Spinoza, is the view that all organization of things in the world is necessary, but “unintentional (because it is derived from an original being, but not from its understanding, hence not from any intention on its part)” (CPJ, §72. 5: 391–392). Both of these doctrines simply explain away any appearance of purposiveness. The realism of purposiveness does not, however. This doctrine also takes two forms, “either physical or hyperphysical.” The first, which Kant calls “hylozoism,” “bases ends in nature on the analogue of a faculty acting in accordance with an intention, the life of matter (in it, or also through an animating inner principle, a world-soul)”; the second, what Kant calls “theism,” derives purposiveness in nature “from the original ground of the world-whole, as an intentionally productive (originally living) intelligent being” which is not itself part of nature (5: 392). Kant then argues that while the two forms of the idealism of purposiveness do no justice at all to our need to comprehend organisms as purposive, hylozoism, the first form of the realism of purposiveness, fails because “the possibility of a living matter…cannot even be conceived.” This is because “lifelessness, inertia, constitutes” the “essential characteristic” of matter (CPJ, §73, 5: 394). In other words, quite apart from the difficulty we might have in conceiving of any particular organic process in purely mechanical terms, there is a more general problem in conceiving of something as living, and thus presumably as having an internal source of change and development,

88

P. GUYER

while also conceiving of it as subject to the law of inertia, thus as capable of changing only when acted upon by something external to itself. For this reason, then, Kant concludes that our only chance for even coherently conceiving of the purposiveness of organisms is by conceiving of their organization as due to the “understanding” and “intentional causality” of an “original being” who lies outside of material nature altogether (5: 395). What makes all of this confusing is, of course, that Kant is arguing for the realism of purposiveness, but only from a reflective and regulative point of view. We might say that he employs the transcendental idealism of purposiveness in order to save its empirical realism. Once we have been brought this far, however, we can see that we can maintain the thought of anything in nature or even the whole of nature as purposive without denying that there is anything within nature that must be essentially resistant to mechanical explanation, because we can see an extramundane ground of nature as accomplishing its intelligent purposes through its design of the mechanical laws of nature. Kant states this point clearly: But since it is still at least possible to consider the material world as a mere appearance, and to conceive of something as a thing in itself (which is not an appearance) as substratum, and to correlate with this a corresponding intellectual intuition (even if it is not ours), there would then be a supersensible real ground for nature, although it is unknowable to us, to which we ourselves belong, and in which that which is necessary in it as object of the senses can be considered in accordance with mechanical laws, while the agreement and unity of the particular laws and corresponding forms, which in regard to the mechanical laws we must judge as contingent, can at the same time be considered in it, as object of reason (indeed the whole of nature as a system) in accordance with teleological laws, and the material world would thus be judged in accordance with two kinds of principles, without the mechanical mode of explanation being excluded by the teleological mode, as if they contradicted each other. (CPJ, §77, 5: 409)

Here we no longer have merely two maxims for explanation, but rather conceptions of two different objects, sensible nature on the one hand and its supersensible ground on the other, which we can conceive of in two different ways, although of course our conception of the latter as intelligent and purposive does not amount to theoretical knowledge. And once we have located real purposiveness altogether outside of nature in this way, we no longer have to insist upon a fundamental distinction between the organic and the nonorganic in order to make room for purposiveness; we could think of organisms too as part of the general system of nature through which the extramundane ground of nature achieves its purposes, even if we cannot understand the laws at work in that part of nature as well as we understand the laws at work in inorganic nature. There might be something special about the experience of organisms and our difficulty in understanding them that leads us to think of anything in nature as purposive in the first place—as Kant puts it, our experience of organisms might be the “particular experiences that bring reason into play in order to conduct the judging of corporeal nature and its laws in accordance with a special principle” (CPJ, §70, 5: 386). But once we have fully thought through what it takes to conceive even of organisms as purposive, we realize that we do not actually have to conceive of them as irremediably distinct from other things in nature in order to conceive of them or anything else in nature as purposive. The claim of the irremediable distinctiveness of organisms becomes a ladder that can be left behind once we have ascended to the more complex view

NATURAL ENDS AND THE END OF NATURE

89

of nature as potentially both fully lawlike and yet the purposive product of an intelligent ground, although Kant hardly makes this explicit. 2.3. Third stage The final third stage of Kant’s argument is the exhibition of the heuristic and regulative benefits of the conception of nature and things within it as purposive. We can conceive of this stage of the argument as itself being introduced in three steps over the course of Kant’s exposition. (a) Kant exhibits the first heuristic implications of the concept of purposiveness in the introductions to the whole third Critique. The published version of the introduction stresses that the “logical” supposition that our concepts of nature’s kinds and laws are systematic and the accompanying “transcendental” supposition that nature itself is systematic are required above all to explain how we can see particular laws of nature as necessarily true even though they are not logically entailed by the most general laws of nature furnished by the pure concepts of the understanding (CPJ, Introduction V, 5: 183). In the first draft of the introduction, however, Kant had suggested another point, namely that the supposition that our concepts of nature form a system actually gives us guidance in the work of reflecting judgment, which is to discover empirical concepts of natural kinds and laws lying between the general categories of the understanding and empirical intuition. Kant says here that without a general principle that “even with regard to its empirical laws nature has observed a certain economy suitable to our power of judgment and a uniformity that we can grasp” (FI, Section V, 20: 213), “all reflection would become arbitrary and blind” (20: 212), thus “we could not hope to find our way in a labyrinth of possible empirical particular laws” (20: 214). He does not spell out how the supposition of the systematicity of empirical laws of nature is supposed to help us find our way about in such a labyrinth. One could suggest that a procedure of first testing empirical hypotheses that are systematically connected with concepts and laws one has already accepted rather than others that might be equally compatible with current empirical evidence would be just the sort of method that would prevent one’s empirical investigations from becoming “arbitrary and blind,” floundering in a labyrinth of possibilities. Of course, such a procedure would come with no guarantee that one’s favored hypothesis might not be upset by subsequent empirical data, in which case the damage that would then be done to the particular hypothesis being tested might also be transmitted into other regions of the system to which it is linked; but at least one would have a method for the conduct of empirical inquiry. Perhaps such a method could also be considered particularly honest precisely because its insistence on systematicity would make one’s whole body of beliefs more rather than less sensitive to new empirical data. (b) The next step in Kant’s exhibition of the heuristic benefits of our conception of purposive nature comes after his argument that we must conceive of organisms as natural ends. He introduces this phase of his argument thus: The concept of a thing as in itself a natural end is therefore not a constitutive concept of the understanding or of reason, but it can still be a regulative concept for the reflecting power of judgment, for guiding research into objects of this kind and

90

P. GUYER thinking over their highest ground in accordance with ends; not, of course, for the sake of knowledge of nature or of its original ground, but rather for the sake of the very same practical faculty of reason in us in analogy with which we consider the cause of that purposiveness. (CPJ §65, 5: 375–376)

This makes it sound as if the conception of organisms as internally purposive natural ends has a pay-off only for practical reason. At the third stage of this phase of his argument Kant will claim that this conception has a pay-off for practical reason, but he will argue first that it has a pay-off for the conduct of scientific inquiry, even if it is not constitutive. However, Kant’s suggestions about this scientific payoff are complex, and again may imply that the distinction between the organic and the inorganic is not insuperable. The first thing that Kant suggests is that the conception of organisms as natural ends gives us a basis for assuming that nothing in them is idle or contrapurposive, and thus for investigating every aspect of them with the expectation that we can find its purpose. Thus he begins the next section by stating the principle that “An organized product of nature is that in which everything is an end and reciprocally a means as well” or that “Nothing” in such an organism “is in vain, purposeless, or to be ascribed to a blind mechanism of nature” (CPJ, §66, 5: 376). This principle is then to encourage us to look for the purposes of even those parts of organisms, such as “skin, hair, and bones,” which might seem to be products of merely mechanical processes without any obvious purpose (5: 377). In the following section, however, Kant seems to make a stronger claim, namely that the conception of organisms as natural ends is actually meant to encourage us to look for mechanical explanations of those processes that we isolate by thinking of their function. Thus he says that the teleological conception of organisms “is not meant to introduce any special ground for causality,” and that the principle of “final causes” is “without harm to the mechanism of nature” (CPJ, §67, 5: 379). Here Kant seems to be continuing a train of thought he had begun in the first draft of the introduction, when he wrote that: E.g., by saying that the crystalline lens in the eye has the end of reuniting, by means of a second refraction of the light rays, the rays emanating from one point at one point on the retina, one says only that the representation of an end is conceived in the production of the eye because such an idea serves as a principle for guiding the investigation of the eye as far as the part that has been mentioned is concerned, with regard to the means that one can think up to promote that effect. No cause acting in accordance with the representation of purposes, i.e., no intentionally acting cause, is thereby attributed to nature.… (FI, Section IX, 20: 236)

And then Kant argues that we cannot place any specific limits on our possible success in the mechanical explanation of organic processes, although we supposedly know in some general way that we will never be able to explain everything in them completely: “we also do not know how far the mechanical mode of explanation that is possible for us will extend, but are only certain of this much, namely, that no matter how far we ever get with it, it will still always be inadequate for things that we once acknowledge as natural ends” (CPJ, §78, 5: 415). This makes it sound as if there is no specific organic process that we can know a priori must resist mechanical explanation, although for some reason we do know that there must be something or other in organisms that mechanical explanation cannot reach. Perhaps Kant’s transition from a model of two incompatible kinds of

NATURAL ENDS AND THE END OF NATURE

91

explanation within nature to a two-level perspective on nature in the course of his solution to the antinomy of teleological judgment is precisely what is required to prevent his conception of the heuristic value of our conception of organisms from collapsing into incoherence. (c) The final step in Kant’s argument, which comes after the resolution of the antinomy of teleological judgment has shown how we can coherently conceive not only of organisms but of the whole of nature as purposive, is the claim that if nature as a whole is a system, then we must conceive of it as if it has some unique purpose. Kant begins by conceding that it is hardly obvious that there need be a single “ultimate end” of nature, that is, a single end to which the natural chain of causes and effects inevitably tends, since it seems indeterminate whether any particular thing in nature should be seen as means or end: to be sure, the existence of human beings can be seen as the end served by the plant kingdom and the herbivores, but then again perhaps humans and the herbivores are merely the means to keep the plant kingdom in proper balance (CPJ, §82, 5: 427). Further, it is certainly not obvious that what many might have supposed to be the ultimate end of nature, namely, human happiness, is an end of nature at all: humans are no more exempt from the “destructive effects” of nature such as “pestilence, hunger, danger of flood, cold, attacks by other animals,” and so on, then anything else (CPJ, §83, 5: 430). Nevertheless, Kant claims, if we are to think of nature as the product of an intentionally acting cause, then we must also be able to think of a purpose for its creation, something of genuine value to be realized through this action: But if we assume that the connection to ends in the world is real and assume that there is a special kind of causality for it, namely that of an intentionally acting cause, then we cannot stop at the question why things in the world (organized beings) have this or that form, or are placed by nature in relation to this or that other thing; rather, once an understanding has been conceived that must be regarded as the cause of the possibility of such forms as they are really found in things, then we must also raise the question of the objective ground that could have determined this productive understanding to an effect of this sort, which is then the final end for which such things exist. (CPJ, §84, 5: 434–435)

This characterizes the structure of our own thought: an intelligent or rational ground is necessary for us to explain the purposive form of the world, but we cannot conceive of an intelligent ground that acts without a reason, a conception of the value and indeed the unconditional value of what it produces; thus we must be able to posit a final end for the creation of nature as a purposive whole. This is not an argument about how nature actually is. The final step in Kant’s argument is then to argue that the only thing we can conceive of as a final end is humanity itself, although only because of the unconditional value of our capacity for freedom and thus our capacity for the achievement of morality: Now we have in the world only a single sort of being whose causality is teleological, i.e., aimed at ends and yet at the same time so constituted that the law in accordance with which they have to determine ends is represented by themselves as unconditioned…The being of this sort is the human being, though considered as noumenon: the only natural being in which we can nevertheless cognize, on the basis of its own constitution, a supersensible faculty (freedom) and even the law of the causality

92

P. GUYER together with the object that it can set for itself as the highest end (the highest good in the world). Now of the human being…,as a moral being, it cannot be further asked why (quem in finem) it exists. His existence contains the highest end itself, to which, so far as he is capable, he can subject the whole of nature, or against which at least he need not hold himself to be subjected by any influence from nature. (CPJ, §84, 5: 435)

In the previous section, Kant had begun to suggest his conclusion by arguing that it is the “culture of training” or “discipline,” that is, the “liberation of the will from the despotism of desires,” rather than the “cultivation of skill,” which is merely the development of means to happiness, that is a candidate for the ultimate end of nature. This makes it sound as if it is the development of virtue alone that we can conceive of as the end of nature. The present passage, however, alludes to Kant’s view that the complete development of human morality includes the highest good and thus not only virtue but also human happiness, not for its own sake but as the product of human virtue. Thus while human happiness is initially excluded as a possible final end for nature, in the dialectical structure that is so common in the arguments of the third Critique it ultimately returns as part of morality as the object of nature. What is the intended benefit of this moral vision of the system of nature? It is, I believe, two-fold. On the one hand, a chain of thought that begins in our experience of the natural world confirms us in a belief that is also mandated by morality itself. In his treatment of the highest good in the second Critique, Kant had argued that the complete object of our moral efforts must include the realization of human happiness as the product of human virtue, but that we can only conceive of nature as a realm in which happiness can be expected to follow from virtue if we postulate, as a matter of pure practical reason, a common author of the laws of both morality and nature—“the existence of a cause of all nature, distinct from nature, which contains the ground of this connection, namely of the exact correspondence of happiness with morality.”20 Now Kant has argued that a chain of thought that begins with a scientific challenge, that of understanding organisms, leads us to a vision of nature as a whole as a system that can only have the development of human morality as its final end, and in which, of course, the realization of this end must therefore be possible. This can only confirm us in our commitment to morality—not, as Kant is careful to insist in his most careful discussions of the highest good,21 by providing us with any incentive for being moral other than pure respect for the moral law, but rather by persuading us that the object of morality can be realized and thereby preventing our commitment to the moral law from being undermined.22 As Kant argues in his famous 1793 essay, we must be able to believe “that everything in moral philosophy that is correct for theory 20

21 22

Critique of Practical Reason, 5: 125; in Immanuel Kant, Practical Philosophy, translated and edited by Mary J. Gregor (Cambridge: Cambridge University Press, 1996), p. 240. See especially the first section of the essay on “Theory and practice,” 8: 278–289. For further discussion of my approach to the highest good, see “From a Practical Point of View: Kant’s Conception of a Postulate of Pure Practical Reason,” in my Kant on Freedom, Law, and Happiness (Cambridge: Cambridge University Press, 2000), pp. 333–371, and “Ends of Reason and Ends of Nature: The Place of Teleology in Kant’s Ethics,” Journal of Value Inquiry (Spring, 2002).

NATURAL ENDS AND THE END OF NATURE

93

must also hold for practice.”23 In the first half of the third Critique, the existence of natural beauty is supposed to confirm us in our belief that nature is a suitable arena for the exercise of human freedom; in the second half, this line of thought beginning with our distinctive experience of organisms is supposed to lead to the same result. The second point that Kant’s argument might suggest, although again it is hardly one that he spells out, is that while if we see human morality as the final end of nature that legitimates our tendency to dominate and exploit the rest of nature, that conception in fact legitimates such exploitation only for our morally acceptable ends, and also implies that our interpretation of our morally acceptable ends must be compatible with their realization within nature as a single system. Thinking of the realization of our own morality as an end set by a unitary system of nature constrains our interpretation of the demands of morality to those realizable within such a system. In practice, this would mean that we could not interpret morality to require the denial or destruction of nature within or without our own skins. So, for example, morality as an end of nature could not possibly require the abnegation of all our natural inclinations, although it can certainly require their regulation. I believe that this is consistent with Kant’s considered position on the relation between morality and inclination in texts ranging from his reflections on ethics in the 1770s to the Religion of 1793–1794.24 Likewise, the supposition that morality is an end of nature would be incompatible with any idea that morality could permit or require wanton destruction of natural resources without regard to the ecology of nature as a whole as an arena fit for continuing human habitation. This would be compatible with, although a considerable extension of, Kant’s view in the Metaphysics of Morals that in determining the permissible treatment of nonhuman nature, even though such nature does not itself possess the rights that humanity does, we must always consider the effects of our conduct on our own moral sentiments.25 It would be nice if Kant had developed this second line of thought further, but he didn’t. Several millennia of ascetic practices might have alerted him to the dangers of abusing the name of morality to abnegate nature within ourselves, but he could hardly have foreseen the destructive pressures that modern population levels could place on nature outside ourselves. But rather than trying to extend

23

24

25

For a similar argument that the function of a belief in the realizabilility of the ends set by morality is not to provide motivation for acting morally but rather to prevent our motivation for so doing from being undermined, see Henry E. Allison, Kant’s Theory of Taste (Cambridge: Cambridge University Press, 2001), pp. 229–235. “On the common saying: That may be correct in theory, but it is of no use in practice,” 8: 288; Gregor, p. 289. See especially Religion, 6: 26–30. For more extensive documentation, see my papers “The Form and Matter of the Categorical Imperative,” in Volcker Gerhardt, Rolf-Peter Horstmann, and Ralph Schumacher, eds., Kant und die Berliner Aufklärung: Akten des IX. Internationalen Kant Kongresses (Berlin: Walter de Gruyter 2001), vol. I, pp. 131–150; “Ends of Reason and Ends of Nature: The Place of Teleology in Kant’s Ethics,” Journal of Value Inquiry (Spring, 2002); and “Kant on the Theory and Practice of Autonomy,” in Ellen Frankel Paul, Fred D. Miller, and Jeffrey Paul, eds., Autonomy (Cambridge: Cambridge University Press, forthcoming). Metaphysics of Morals, Doctrine of Virtue, §§ 16–18, 6: 442–444.

94

P. GUYER

Kant’s thought far beyond the limits of his own imagination, I will now conclude by asking what if anything in this elaborate vision I have just described might conceivably withstand contemporary criticism.

3. TELEOLOGY SALVAGED? Here I will consider two obvious objections to Kant’s attempt to reconstruct traditional teleology, and consider what if anything in his vision might survive if these objections are sound. The first objection that will naturally suggest itself is aimed at Kant’s claim that we must presuppose that nature itself is the systematic product of an intelligent design in order to make it rational for us to seek a unique system of empirical laws that can in some sense be seen as necessary truths. One part of this objection is simply that we may now assume neither that an adequate theory of nature must be unique nor that it makes sense to think of any adequate theory of nature as necessarily true. Thus from our point of view there may be no need to render comprehensible that which Kant thinks we must. The other part of this objection is that we may not need a guarantee of success in finding systematicity in order for the pursuit of it to be rational, as long as we have no reason to think that success is impossible. I think that the reply to both of these points will be that, regardless of what the postmodern philosopher may say, the working scientist is constantly seeking to integrate his explanation of any particular realm of nature into a single larger explanatory framework. He will hardly regard success in that venture as guaranteed, but neither could he imagine why he should not try. The scientist might well be indifferent to the philosopher’s question whether the laws of nature are necessary or contingent, but will assume that the uniqueness and comprehensiveness of the system of empirical laws are the natural goals of science. Presumably Kant’s supposition that the preference for systematic rather than nonsystematic hypotheses offers at least a methodology or research strategy for empirical inquiry is tacitly assumed by every working scientist. The second objection will be to Kant’s claim that the reciprocal causation involved in organic processes defies ordinary mechanical explanation, and forces us to conceive of organisms as products of intelligent and purposive design instead. The whole progress of biological science since the middle of the 19th century, it will be argued, ranging from biochemistry, genetics, the theory of natural selection, and down to contemporary genomics, is to show how we can give mechanical explanation of the complex functions and appearances of design that we find in organisms. We have no reason to believe that there is anything in organisms that necessarily defies mechanical explanation. As already suggested, Kant himself clearly worried that any particular limitation on mechanical explanation might be unjustified. And thus he admitted, for example, that The agreement of so many genera of animals in a certain common schema, which seems to lie at the basis not only of their skeletal structure but also of the arrangement of their other parts, and by which a remarkable simplicity of basic design has

NATURAL ENDS AND THE END OF NATURE

95

been able to produce such a great variety of species by the shortening of one part and the elongation of another, by the involution of this part and the evolution of another, allows the mind at least a weak ray of hope that something may be accomplished here with the principle of mechanism, without which there can be no natural science at all. (CPJ, §80, 5: 418)

Here he fell back on a vague general claim that the possibility of reproduction itself would presuppose “an organization purposively aimed at all these creatures” (5: 419), even if the variety of their particular forms could be explained by what we would now consider to be the evolutionary mechanisms to which he alludes. But he obviously did not feel confident that even a general claim that the possibility of life or reproduction itself could not be mechanically explained would withstand critical scrutiny. As we have seen, Kant’s solution to the antinomy of teleological judgment ultimately leaves this claim behind. This issue was also a major concern for Kant in his uncompleted final work, what we call the Opus postumum, where he argued on the one hand that even matter itself must have an internal source of motion, in the form of the ether, thus putting into question his earlier claim that matter cannot be living because of its inertia, while on the other he sought to characterize the distinction between the organic and the inorganic as the first distinction to be drawn within a system of the physical forces of material nature.26 So in the end Kant himself was at best undecided about the justification for any claim that we must necessarily comprehend the organic and the inorganic in two different ways. But does that mean that Kant’s analysis of the heuristic benefits of our conception of organisms as purposive must also come to naught? I do not think so. Kant argued that conceiving of organisms as the products of purposive design has two benefits for us: it lead us to seek for the explanation, ultimately the mechanical explanation, of the functions that we attribute to them; and it leads us to seek for a purpose for their creation and the creation of all of nature, which we can only find in our own moral development, but in our moral development within the system of nature. He could still find these benefits in our conception of organisms as purposive even if he held only that such a conception of organisms is merely suggested to us by the phenomenology of our experience of organisms. That is, he could argue that the teleological point of view is natural to human psychology, and then, on his general teleological assumption that everything that is natural should be assumed to have a proper purpose, find the purpose of this tendency in human psychology in its usefulness to scientific inquiry on the one hand and its value for our commitment to morality on the other. He could then argue, as he argues in the Metaphysics of Morals about our natural dispositions in behalf of the beauty of the inorganic and the nonhuman, that even though such dispositions are not absolutely necessary to either scientific inquiry or moral conduct, they should nevertheless be preserved and promoted because they are beneficial to our pursuit of both science and morality in the empirical conditions of real life.

26

For further discussion of these issues, see my “Organisms and the Unity of Science,” in Watkins, ed., Kant and the Sciences.

96

P. GUYER

This is how I understand Kant’s claim that a human being has a duty to himself to refrain from whatever “weakens and gradually uproots a natural predisposition that is very serviceable to morality.”27 Our conception of organisms as purposive can be thought of as a natural predisposition that is serviceable to both science and morality, and which should be preserved rather than weakened and uprooted simply for that reason. Kant’s rescue of teleology could thus be seen as building upon but completing Hume’s own mitigated scepticism toward it. Hume, through Philo, ultimately admitted that it is a natural tendency of the human mind to posit at least some sort of cause for the whole world he experiences. Kant goes well beyond Hume in diagnosing the illegitimacy of such a tendency as a pretension to theoretical insight. But he also shows that as a piece of empirical psychology, Hume’s thought is incomplete: if we do naturally incline to think of organisms and the world as a whole as purposive, we equally well think of them as having a point. And once we see that we can use this thought to serve both our scientific and our moral progress, we will realize that we have no reason to suppress it, but every reason to let it serve in this way. University of Pennsylvania

27

Metaphysics of Morals, Doctrine of Virtue, § 17, 6: 443; Gregor, p. 564.

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED KELD NIELSEN AND HANNE ANDERSEN

1. INTRODUCTION In attempting to answer the question why it was H. C. Ørsted (1777–1851) who discovered electromagnetism, historians of science have discussed whether Ørsted was influenced by Schelling’s Naturphilosophie or rather by Kant’s philosophy. Thus, R. C. Stauffer and L. P. Williams have seen Ørsted’s discovery of electromagnetism as closely related to the influence which J. W. F. Schelling’s Naturphilosophie had exerted on Ørsted prior to the discovery; the main point being that Ørsted did not perceive of electricity, heat, light, and magnetism as imponderable, corpuscular fluids but as forces of a common origin.1 A contrary view has been put forward by Shanahan, who claims that the most important background to Ørsted’s discovery was his knowledge of Immanuel Kant’s philosophy.2 Although Shanahan agrees with Stauffer and Pearce Williams in that Ørsted’s ideas of forces were essential to the discovery, he claims that the decisive influence came from Kant’s Metaphysical Foundations of Natural Science published in 1786. According to Shanahan, this book introduced Ørsted to ideas that were to constitute the core of his scientific thinking throughout his life, and, furthermore, that Kant’s insistence on the limited applicability of a priori methods in science contained a vital antidote to Schelling’s rambling speculations. In Shanahan’s words “it was [Ørsted’s] deeply rooted acceptance of certain Kantian doctrines, and his explicit rejection of central doctrines of [Schelling’s] Naturphilosophie, that made this discovery possible.”3 However, as Caneva has later pointed out, “no one has yet studied Ørsted’s work in its totality with due attention to the full range of his philosophical and scientific sources.”4 The purpose of this paper is to undertake only a small part of this overwhelming task by examining in detail to what extent and in which way the 1

2

3 4

R. C. Stauffer, “Persistent Errors Concerning H. S. Ørsted’s Discovery of Electromagnetism,” Isis 44 (1953), pp. 33–50; L. P. Williams, The Origin of Field Theory (New York: Random House, 1966). T. Shanahan, “Kant, Naturphilosophie, and Ørsted’s discoversy of electromagnetism: A reassessment,” Studies in the History and Philosophy of Science 20 (1989), 287–305. Shanahan, “Kant, Naturphilosophie, and Ørsted’s discovery…, p. 289. K. Caneva, “Physics and Naturphilosophie: A Reconnaissance,” in Hist. Sci 35 (1997), pp. 35–106.

97 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 97–114. © 2007 Springer.

98

K. NIELSEN AND H. ANDERSEN

young Ørsted was influenced by Kantian philosophy. Our investigation deals with the period prior to 1802, the year in which Ørsted set out for his great educational journey to Germany and France. First, we shall discuss whether it is possible to point to specific factors which inspired Ørsted to take up Kant. Next, we shall discuss in what way and to what extend Ørsted accepted certain Kantian doctrines. Finally, we shall discuss how this investigation of Ørsted’s writings on Kant may affect the existing views on the influence of Kant’s philosophy on Ørsted.

2. ØRSTED’S PRE-KANTIAN PERIOD 1794–1797: THE PROMISING AND DILIGENT STUDENT Hans Christian Ørsted was born in 1777 in Rudkjøbing, the largest town on the small island of Langeland some 100 km from Copenhagen. Hans Christian and his two years younger brother Anders Sandøe both showed promising intellectual abilities, and in 1794 they together moved to Copenhagen to prepare for the entrance examination for the university. The philosophical examination was passed by both of them in 1795 with distinction. At this time Hans Christian was very fond of poetry and wrote several poems to develop his skill, but he was also interested in astronomy, mathematics, and physics, and he decided to study medicine. Anders Sandøe was very interested in philosophy, but decided to study law.5 During the winter of 1795/96 Hans Christian Ørsted left Copenhagen for a while to go back to his native town to practice laboratory work under the guidance of his father who was a pharmacist with his own pharmacy and a very active chemist. In April 1797 his father certified that Ørsted possessed the skills required by a professional pharmacist, and that Ørsted had been present during the most elaborate experiments he had carried out.6 Thus, there is no doubt about the origin of Ørsted’s experimental interest and skills. Although Shanahan in his description of the background to Ørsted’s discovery of electromagnetism emphasizes how Kant repeatedly mentioned the need for experiments and observations and hence sees the writings of Kant as the principal reason why Ørsted performed experiments, he overlooks that Ørsted was brought up in a laboratory and not in a library.7 This is not to say that Ørsted was not an ardent reader, but his interest in experiments seems to have been developed early in his career, long before he took interest in Kant. In 1796 the University of Copenhagen announced an essay contest within aesthetics on the topic, How can the prose language be corrupted by coming the poetic too near? and where are the limits between the poetic and the prose expression? Ørsted submitted an essay in which the main emphasis was on the position of A.G. Baumgarten (1714–1762), while there were no traces of neither romanticism, nor of Kant. As Ørsted wrote, he had not included what more recent philosophers had said about the nature of poetry since he had found no author to be more 5 6

7

Anders Sandøe Ørsted was later to become a highly influential and respected civil servant. K. Baerentsen and V. G. Jensen, “Om den unge Ørsted ,” Theriaca. Samlinger til farmaciens og medicinens historie (Copenhagen: Dansk Farmacihistorisk Selskab, 1977), pp. 11–21. Shanahan, “Kant, Naturphilosophie, and Ørsted’s discovery…,” p. 304.

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

99

profound than Baumgarten.8 Ørsted won the gold medal, and his essay was published in the journal Minerva in May 1797. Also in May 1797, Ørsted passed the final examination in pharmacy with distinction. One of the examiners, the professor of medicine J. C. Tode, found the knowledge exhibited by the student so impressive that he gave a whole page in his own journal to the description of this happy incident.9 At the beginning of 1797, the medical faculty at the University of Copenhagen announced an essay contest on the origin and function of the amniotic fluid. Ørsted had in his first prize essay showed his ability to orient himself in a new field, developing an independent position, and he also possessed both the knowledge of chemistry and the experimental experience necessary to participate in the contest, and he submitted his second prize essay before the end of the year. The essay was well researched. He had references to such authorities as Blumenbach, von Haller, Van den Bosch, Malpighi, Wünsch, Buffon, and several less-known authors. His own chemical investigations were all qualitative tests in which he tried to nail down the components of the amniotic fluid. His main conclusion was that the fluid mainly consists of a thin, salty solution of proteins. Rhetorically, the paper was held in style of the inductive tradition so common at the time. It contained several warnings against conclusions not based on careful observations. Thus, in the introduction the young author warned against the seductive workings of the inquiring human mind: With the bright colours of imagination it paints what reliable experience has not shown it and creates easily overthrown theories where facts do not guide it. However, if we do not want to give up for ever the hope of attaining some certainty in this, we must first investigate more closely the part of it which immediately offers itself to our senses, for the disagreement of the greatest natural philosophers prove that even in this we have not proceeded as far as we could have.10

This view returned again later in the essay, when Ørsted had to admit that the sum of knowledge about the origin of the amniotic fluid was insufficient to decide conclusively on the question of the origin of the amniotic fluid. Here Ørsted stated that I do not think that we know anything more definite about the origin of the amniotic fluid at the present time. Perhaps some fortunate discoveries of the future will be able to shed more light on this issue. We must only try as much as possible to avoid theories about it as long as we do not have sufficient data on which to build them.11

There was not a single trace of Kant’s philosophy in this essay either, but clear evidence of a young scholar eager to convince the professors of the university that he was playing strictly by the recognized rules, that only when sufficient facts are at hand theories may be formulated—cautiously. The official evaluation of 8

9 10

11

F. J. Billeskov Janssen, “Hans Christian Ørsted,” in F. J. Billeskov Jansen, Eigeill Snorrason and Chr. Lauritz-Jensen (eds.), Hans Christian Ørsted (Hellerup: IFV-energi i/s, 1987), p. 15. Bærentsen and Jensen, “Om den unge Ørsted…,” p. 17. Hans Christian Ørsted, “Om Modervandets Oprindelse og Nytte”(1797), translated as “Response to the Prize Question in Medicine Set by the University of Copenhagen in the Year 1797: On the Origin and Use of the Amniotic Fluid,” in K. Jelved, A. D. Jackson and O. Knudsen (eds. and trans.), Selected Scientific Writings of Hans Christian Ørsted (Princeton, NJ: Princeton University Press, 1998), 4., pp. 3–25. Ibid p. 14.

100

K. NIELSEN AND H. ANDERSEN

Ørsted’s prize essay stressed three points: the essay was well organized with clear and thorough descriptions; in the chemical analysis Ørsted had shown diligence; and he had tried to mediate the conflicting views of chemistry of that time.12 To temper claims about Kant’s influence in relation to Ørsted’s understanding of the importance of experimental evidence, it should also be noted that Anja Jacobsen has convincingly argued that in his methodological novelty in this prize essay—the use of chemical analysis to decide a medical question—Ørsted was inspired by the view of chemistry introduced by Lavoisier.13 Again, Ørsted won the gold medal. Hence, by now he had distinguished himself three times on quite different subjects in the small academic circles in Copenhagen: by the prize essay in aesthetics, by the prize essay in medicine, and by passing the final examinations in pharmacy with distinctions. He had fitted perfectly into the role of the well-behaved prodigy who played strictly by the recognized rules.

3. THE SPREAD OF KANTIAN PHILOSOPHY During the last two decades of the 18th century, Kant’s philosophy spread from Königsberg through Germany, and gradually to other European countries as well, such as Holland, Sweden, and Denmark.14 Kantian philosophy was in opposition to the reigning Leibniz-Wolffian philosophy, and it met much resistance, mainly due to its various political and theological implications. In many places, the conflict took form of a generational clash. In Stockholm, for example, eager young Kantians stated publicly that the old philosophical system was in need of reform and that such a reform had been initiated and carried out by the genius Kant, while their main opponent, professor Christiernin, had attacked “the new philosophy” in the newpaper Uppsala Tidning.15 Feelings were so intense that it came to a public confrontation at which—so the story goes— Christiernin opposed the gathering youth, giving the academic guards orders to “beat them, stab them, shoot them, I take full responsibility for any loss of life.” After this incident the government had to intervene, and Christiernin was suspended by the university’s vice-chancellorship.16 In Copenhagen, Kant’s philosophy became immensely popular among young people during the later half of the 1790s.17 Among the most eager young Kantians

12

Ibid p. 3f. A. S. Jacobsen, Between Naturphilosophie and Tradition. Hans Christian Ørsted’s Dynamical Chemistry (Unpublished Ph.D. Thesis, University of Aarhus, 2000), pp. 8–9. 14 Van der Wyck, “Kant in Holland,” Kant-Studien 3 and 8 (1899 and 1903), pp. 403–414 and 448–466; M. R. Wielma, “Die erste niederlaendische Kant-Rezeption 1786–1850,” Kant-Studien 79 (1988), pp. 450–466; A. Vannerus, “Der Kantianismus in Schweden,” Kant-Studien 6 (1901), p. 244–269; A. Thuborg, Den kantianske periode dansk filosofi 1790–1800 (Copenhagen: Gyldendal, 1951). 15 Vannerus, “Der Kantianismus in Schweden…,” p. 260. 16 Ibid p. 262. 17 For an overview of the early Kantianism in Denmark, see Thuborg, Den kantianske periode i dansk filsofi…(n. 15). 13

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

101

was Hans Christian Ørsted’s brother Anders Sandøe Ørsted, with whom Hans Christian shared room and money until the summer of 1801. In Denmark, the first reference in writing to Kant seems to be an anonymous article from 1789 published in the periodical Minerva with the title: “The Kantian definition of thought.” The first university lectures on Kant’s philosophy were held in 1793 by C. Hornemann who had been studying Kant’s writings since 1784.18 In 1791 Hornemann had gone to Jena to hear Reinhold, Fichte’s predecessor, and upon his return he started a series of lectures on Critique of pure Reason, but he died prematurely at the end of 1793. Hornemann’s lectures were popular, and they invoked a remarkable Kant-enthusiasm among young intellectuals in Copenhagen. An intense debate developed in a number of periodicals, and around 1796 a downright craze had developed. Silver medals showing Kant’s portrait were sold, it was suggested to make Kant’s philosophy the basis of politics and of the judicature system, Kant was claimed to be the new Messiah, etc. Anders Sandøe entered the public debate in 1797, but already the year before he had found reading Kant to be the best cure against weak nerves. In 1798 he portrayed Kant as a giant who had managed to crush all earlier philosophical systems and stop their destruction of moral order. To Anders Sandøe, Kant had given the moral sciences a new foundation that “would defy eternity.”19 3.1. The prudent advocate Ørsted began making references to Kant’s philosophy in 1798. During 1798 and 1799 he worked on three projects more or less simultaneously: A series of four essays on chemistry, a critical essay review of a new edition of the leading Danish textbook on physics and chemistry, and a long essay on Kant’s metaphysics of science. In these three projects he, with increasing enthusiasm, presented Kant’s “dynamical system” as a superior alternative to the reigning atomism. The four essays on chemistry were published under the title of “Letters on Chemistry” in the same periodical which had brought his prize essay on the amniotic fluid.20 In form and content the letters were clearly aimed at a lay audience. The main subject was heat and combustion and in a plain language Ørsted provided pedestrian explanations of various fundamental ideas: the difference between physics and chemistry, the function of a thermometer, the process of distillation, the meaning of concepts like latent heat, heat capacity, heat conductivity, etc. Without any reservations, combustion was explained the basis of the sanctioned theory of Lavoisier.

18 19 20

Ibid p. 10. Ibid p. 17. Hans Christian Ørsted, “Letters on Chemistry. First Letter (1798),” “Letters on Chemistry. Second Letter, on Heat (1798),” “Letters on Chemistry. Third Letter (1799),” “Letters on Chemistry. Fourth Letter (1799),” in Jelved, Jackson, and Knudsen, Selected Scientific Works of Hans Christian Ørsted…, pp. 26–28; 29–34; 35–40; 41–45.

102

K. NIELSEN AND H. ANDERSEN

Kant was briefly mentioned in the first of the letters, in which Ørsted claimed that matter possesses two powers, a cohesive force and an expansive force. He emphasized that he could “not, without straying too far from [his] goal, inform [the reader] of the profound investigations on which depends the evidence that bodies really do possess these fundamental forces,” but instead limited himself to mentioning some experiences which, to some extent, could serve to convince [the reader] of the existence of the cohesive force at any rate. If you take two polished glass plates and put one on top of the other, you observe that they stick together; if you put one drop of mercury next to another, they unite; and if you dip your finger into water, the moisture sticks to it; all this proves that one body displays a cohesive force with another.21 One may here note the remarkable difference between this rather thin experimental evidence for the existence of the two fundamental forces and his warnings the previous year against letting the human mind create “easily overthrown theories where facts do not guide it.” The second chemical letter dealt specifically with the nature of heat. Like Kant, Ørsted claimed heat to be material. In a note he added that he had always considered heat to be material since it could pass from one body to another and a mere quality could not do that. However, he also noted that he had later been further convinced by the argument that “bodies have only two fundamental forces, the expansive force and the cohesive force, and it is possible to derive all others forces from these. The cause of heat acts in such a way that it cannot be conceived as one of these original or derived forces, so it must be material.”22 In his explanation of the various phenomena related to heat, Ørsted was utterly conventional. He operated with the material heat as being “free” or “bound,” and he explained the transition of a body from one state to another as a process in which the body liberates or binds material heat, and thus reproduced the French caloric theory. Although he had started referring to Kant’s dynamical view of matter, he still remained largely within the accepted chemical framework.

4. ØRSTED’S TRANSITORY PERIOD 1798–1799 4.1. The provocative reviewer In his second project from the years 1798–1799, Ørsted went a step further to use Kant’s dynamical theory of matter as his main weapon in an attempt to challenge the socially most prominent scientist in Copenhagen, the Lord Steward, Adam Wilhelm Hauch. Hauch had introduced Lavoisier’s chemistry in Denmark, and he was also the owner of one of the largest collections of scientific instruments in the world, of course by far the largest in Copenhagen. Hauch was head of the royal stables and gave lectures in his “Physical Cabinet.” His textbook on natural science from 1794 became quite popular and the first

21 22

Ørsted, “Letters on Chemistry. First Letter,”…, p. 27. Ibid p. 29n.

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

103

volume of a second, expanded edition appeared in 1798. Ørsted made an essay review in three parts for one of the leading journals, attacking Hauch for not taking Kant’s dynamical view of matter into account. Thus, Ørsted’s main objection against Hauch’s book “concern[ed] the author’s philosophy of nature. He follows here the atomistic system.”23 According to Ørsted “this system leaves the question unanswered: Why do bodies occupy space?” Ørsted’s main argument was that any atomistic answer to the question “Why does matter occupy space?” would be circular: Atomists claim that matter fills space because it consists of impenetrable particles. How can these particles be impenetrable? Because they fill space, etc. Thus, one was led to the expansive force of the dynamic system which the atomic system had intended to avoid. Ørsted mentioned F. A. C. Gren and reproached Hauch for failing to mention how this writer had shown “how the results of the new philosophy could prove useful for physics,” but he also emphasized that he did not want to enter into more detailed arguments since he intended to publish a more detailed survey very soon.24 The publication of the second volume of Hauch’s textbook was delayed, and this gave Hauch the opportunity to add a supplement in which he answered Ørsted’s objections. He accepted some of the more trivial points raised, though not all of them, but he did not simply accept the objections to the atomistic system. Although he did admit the dynamical system some ground, he “did not dare to judge the one system better than the other, but was convinced that they both deserve to be known.”25 Hence, Hauch did not give up his position, but instead demonstrated his knowledge of both systems and argued that in the dynamical system the assumed repulsive force was superfluous. In an elegant manner he also revealed that he knew the identity of the anonymous reviewer by mentioning that this reviewer as well as Mr. Ørsted were defenders of the dynamical system, which hitherto “had not been explained with such clarity as one could wish” and held his arguments against the dynamical system as well as the rest of his response in a polite tone.26 Ørsted stated his appreciation of this in the second part of his review, which was now no longer anonymous.27 In this second part of the review Ørsted again stressed the advantages of the dynamical system, but now in an even more daring manner. He might have hoped for a fight, but what he had obtained was even better: Hauch had treated him as a peer. In July 1800, a chair in physics at the University of Copenhagen became vacant. Ørsted wanted the chair but knew that the influential professor of astronomy, Thomas Bugge, would be against his candidacy. To get support Ørsted wrote to

23

24

25

26 27

H. C. Ørsted, “Begyndelesgrunde til Naturlaeren. Anden og forbedrede Udgave ved A. W. Hauch. Anden Del (review of Foundations of Physics, 2nd ed., by A. W. Hauch, Part Two), Kjøbenhavns laerde Efterretninger, reprinted in H. C. Ørsted, Naturvidenskabgelige Skrifter (Scientific Papers), 3 vols., edited by K. Meyer (Copenhagen: Høst og Søn, 1920), vol. III, p. 28. Ibid. p. 29. Ørsted here alluded to his Fundamentals of the metaphysics of nature which will be dealt with below. Hauch, Begyndelsesgrunde til Naturlaeren. Anden og forgbedrede Udgave, Anden Del (Foundations of Physics, 2nd revised ed., part 2) (Copenhagen, 1799), p. 770. Ibid. Ørsted, “Begndelsegrunde til Naturlaeren…,” p. 38.

104

K. NIELSEN AND H. ANDERSEN

P. C. Abildgaard, the founder of The Royal Veterinary College, who answered that he had already told a few friends that he would prefer to see Ørsted in that chair. Further, he recommended Ørsted to approach Hauch for advice and support, “for I know that he has much respect for your expertise.”28 Ørsted’s review had not been without effect.

5. KANT’S METAPHYSICAL FOUNDATIONS OF NATURAL SCIENCE The third of Ørsted’s three projects was to introduce Kant’s Metaphysical Foundations of Natural Science to a Danish audience. In Metaphysical Foundations Kant elegantly combined two objectives. On the one hand Kant hoped that a better understanding of his philosophy could be achieved by giving an example in concreto of how to use the abstract principles and concepts of the transcendental philosophy.29 On the other hand, Kant wanted to use the principles of transcendental philosophy to find the a priori sources for the Newtonian laws and concepts and thus provide apodictic certainty to the most prominent contemporary theory which served as a paradigm for how science ought to be.30 Hence, by analyzing Newtonian mechanics according to the principles of transcendental philosophy, Kant obtained both to show how his philosophy worked in actual practice, and to provide Newtonian mechanics with an a priori foundation. The basis for Metaphysical Foundations was thus the principles of Kant’s Critique of pure Reason.31 In this work Kant had described how our knowledge springs from two sources of the mind: the capacity of receiving the yet unprocessed impressions of some object, and the capacity to recognize the object from these representations through the process of thought. When we experience something, an object affects our faculty of representation and gives rise to a sensation. This sensation is transformed into an appearance by the two forms of intuition, time and space. Thus, the appearance is a representation of the object which has not yet been processed by thought. Thought is then involved when a particular representation is recognized as an instance of a concept. The act of recognizing a particular representation as an instance of a concept is called the judgment. The basic concepts of the understanding are the categories. By abstracting from the actual content of all possible judgments, Kant deduced twelve basic forms which he divided into four classes: quantity, quality, relation, and modality. Since on Kant’s view all judgments had to take one of these twelve forms, the 28

29

30

31

Abildgaard to Ørsted, 22, August 1800, in M. Ørsted, Breve til og fra Hans Christian Ørsted (Letters to and from Hans Christian Ørsted), 2 vols., (Copenhagen: Th. Lind, 1870), vol. 1, p. 6. Michael Friedman, Kant and the Exact Sciences (Harvard, MA: Harvard University Press, 1992); P. Plaas, Kants Theorie der Naturwissenschaft (Goettingen, 1965). I. Kant, Die Metaphysische Anfangsgrunde der Natürwissenschaft [1786] (Frankfurt am Main: Suhrkamp Verlag, 1977), p. A: XIII; Friedman, Kant and the Exact Sciences…, section 3.1. The first edition of Critique of Pure Reason appeared in 1781, but in 1786 Kant made a major revision which appeared the following year. In the meantime he had published Prolegomena to Any Future Metaphysics in 1783. Metaphysical Foundations appeared at the same time as he revised Critique of Pure Reason, in 1786.

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

105

basic conditions for any experience can be derived by applying the twelve basic forms of judgments to the concepts involved in the experience in question. Metaphysical Foundations was the concrete application of all these abstract rules on our experience of the external world, that is, on the objects that affect our external senses. For an object to be an object of the external senses, the object must be sensible—“something” must be sensed by us. The aim of Metaphysical Foundations was therefore to show which a priori judgments could be deduced by applying the twelve categories to this “something,” the external matter. But instead of simply applying the twelve categories on the concept “matter,” Kant made an important move. By the end of the introduction to Metaphysical Foundations he claimed that “[t]he fundamental determination of a something that is to be an object of the external senses must be motion for thereby only can these senses be affected. The understanding leads all other predicates which pertain to the nature of matter back to motion; thus natural science is throughout either a pure or an applied doctrine of motion.”32 Kant did not extend his argument to explain why the external senses could only be affected by motion. However, if matter is to be experienced, it must first be represented as an appearance, where the appearance is the sensation ordered in both time and space. Purely spatial appearances would be geometry, and appearances purely in time would be arithmetic, but the combination of time and space is motion.33 Thus, based on the claim that the fundamental determination must be motion, his investigation of the basic conditions for our experience of the outer world was transformed from an investigation of which a priori judgments can be deduced by applying the categories to the concept of “matter” to an investigation of which a priori judgments can be deduced by applying the categories on the concept of “motion.” Therefore, Metaphysical Foundations contained four parts, investigating motion with respect to quantity, quality, relation, and modality. Although Metaphysical Foundations seem to have attracted less attention than some of Kant’s other works, it found plenty of adherents.34 Among the contemporary scientists such figures as von Arnim, Baader, Hildebrandt, Richter, Scherer, and Weiß explicitly adopted it. The methodology of Metaphysical Foundations, to subsume something given under the categories, was broadly adopted among scholars of various fields as “applications of Kant’s philosophy” on the respective field. However, works adopting this methodology were often attacked from two sides: being too little speculative for the philosophers, but to short of scientific facts for the scientists. 32 33

34

Kant, Metaphysische Anfangsgründe…, p. A: xxi. Similar interpretations have been suggested by, e.g. Schäfer 1966 and Plaass 1965. It should be noted that the central claim that motion must be the fundamental determination was treated in different ways by Kant’s contemporary readers. Thus, while some reviewers of Metaphysical Foundations quoted the claim unquestioned (e.g. Wirzburger gelehrte Anzeigen July 28, 1787, Philosophische Annalen April 1787, Allgemeine deutsche Bibliothek June 1787; all in Landau 1991) others critically asked whether the claim was a statement a priori or not (Jenaische gelehrte Anzeigen March 5, 1787, in Landay 1991). However, in this paper we shall not enter the discussion of how this controversial claim of Kant’s is to be interpreted, but only draw attention to the fact that the claim is essential to the whole project of Metaphysical Foundations; a fact that is important when turning to Ørsted’s writings. M. Carrier, “Kants Theorie der Materie und ihre Wirkung auf die zeitgenoessische Chemie,” Kant Studien 81 (1990), pp. 170–210.

106

K. NIELSEN AND H. ANDERSEN 6. ØRSTED’S KANTIAN PERIOD 1799–1802: THE YOUTHFUL, RADICAL KNOW-IT-ALL

Ørsted’s third project from this period—to introduce Kant’s Metaphysical Foundations to a Danish audience—was also the most ambitious. Rather than simply translating the book, Ørsted made a 40-page paraphrase which also followed a different structure than Kant’s book and he therefore titled it Fundamentals of the metaphysics of nature. Partly according to a new plan.35 It was printed in 1799 in the leading pro-Kantian journal Philosophisk Repertorium on which editorial board both the Ørsted brothers served. Later the same year Ørsted reworked Fundamentals of the Metaphysics of Nature into a thoroughly revised and slightly shorter version in Latin which he submitted as his doctoral thesis with the title Dissertation on the Structure of the Elementary Metaphysics of External Nature.36 Finally, in 1802 while Ørsted was in Germany as part of his educational journey, a third, revised version was published in German under the title Ideas of a new architecture of the metaphysics of nature.37 Two problems immediately catch the eyes in Ørsted’s interpretations of Metaphysical Foundations. First, Ørsted discarded Kant’s central claim that the fundamental determination of a something that is to be an object of the external senses must be motion. Second, apparently Ørsted did not understand the a priori character of Kant’s concept of matter, and accused Kant for building his theory on empirical grounds. As described above, Kant claimed that motion was the fundamental determination of a something that is to be an object of the external senses. Consequently, the understanding would trace all judgements of matter back to judgements of motion. However, Ørsted took a completely opposite view and separated the investigation of matter from the investigation of motion. He therefore went through the four classes of categories twice. He did this in each of the three works in which he built on Kant’s Metaphysical Foundations, that is, Fundamentals of the Metaphysics of Nature from 1799 (Fundamentals), Dissertation on the Structure of the Elementary Metaphysics of External Nature likewise from 1799 (Dissertation), and Ideas of a new architecture of the metaphysics of nature from 1802 (Ideas). It is noteworthy that Ørsted did not provide much argument for why he opposed Kant on this very central point. In the first of the three editions, the Danish Fundamentals, he even provided an argument for Kant’s claim that motion is the fundamental determination of a something that is to be an object of the external senses:

35

36

37

Hans Christian Ørsted, “Grundtraekkene af Naturmetaphysikken. Tildeels efter en nye Plan,” reprinted in Ørsted, Naturvidenskabelige Skrifter…, pp. 46–78. H. C. Ørsted, “Dissertation on the Structure of the Elementary Metaphysics of External Nature,” in Jelved, Jackson, and Knudsen, Selected Scientific Writings…, pp. 79–100. H. C. Ørsted, D. Johann Christian Ørsted’s Ideen zu einer neuen Architektonik der Naturmetaphysik. (Berlin: D. W. H. Mendel, 1802).

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

107

Nothing can be perceived by the external senses except in time and space. When we contemplate matter, which is what we call each object for the external senses, in these two forms, the notion of motion and rest arises, the former by contemplating it at different times in changed relations to the rest of space, the latter by contemplating it as present in one space for some time. Here, then, we see the emergence of a pure doctrine of motion.38

However, after having provided this argument explaining why the fundamental determination of anything that has to be an object of the external senses must be motion, he immediately added that: in addition to this, the metaphysics of nature must also contain another part, which will teach us what properties we must necessarily presuppose in matter if it is to become an object of our external senses. This might be called a pure doctrine of matter.39

Consequently, he concluded that each of these two doctrines had to be taken through the four classes of categories, and he therefore ended with a division into eight parts. He realized that this was at odds with Kant’s work and admitted in the introduction that, “[a]nyone who is familiar with Kant’s metaphysics of nature will see that this division is somewhat different from that of the famous thinker.”40 However, as an actual argument for his deviation he merely hoped “that the following will demonstrate sufficient grounds for my deviation. Here I can state no other than the one I have just put forward, namely, the division itself.”41 Ørsted’s opposition to Kant’s central claim developed with each of his three works. In the Danish Fundamentals, he introduced the doctrine of matter by the argument quoted above and treated the doctrine of motion first and the doctrine of matter second. However, in the opening sentence of the treatment of the quantity of matter he stated that As we have seen above, that which is movable in space is called matter, and we regarded it in this manner throughout the doctrine of motion, but the first notion we have of matter is that it is the object for the external senses.42

Apparently, not only did he not accept Kant’s claim that the understanding leads all other predicates which pertain to the nature of matter back to motion, he even insisted that the notion of matter was prior to the notion of motion. This led to a major structural change of his treatment when he revised it into the Latin version, Dissertation, published later the same year. Here he reversed the order of the sections, treating the doctrine of matter before the doctrine of motion, with the argument that The doctrine of matter must come before the doctrine of motion, for without knowledge of extension, shape, and several other properties of matter, no doctrine of motion can be established.43 38 39 40 41 42 43

Ørsted, “Fundamentals of the Metaphysics of Nature…,” p. 47f. Ibid p. 48. Ibid. Ibid. Ibid. p. 59. Ørsted, “Dissertation on the Structure of the Elementary Metaphysics of External Nature…,” p. 83.

108

K. NIELSEN AND H. ANDERSEN

This forced him to introduce two new definitions of matter and motion, since matter could no longer be defined simply as the movable. Instead, he suggested that What is real and is located in space we call matter, and its changes, which likewise can only occur in space, we call motion. Thus, two doctrines arise, one of matter, the other of the motion of matter.44

This seems to be a quite naïve realist view quite at variance with Kant’s position. Thus, on a truly Kantian view, space (and time) only pertain to the appearances which arise when the two forms of intuition transform sensations, but not to the objects that give rise to the sensations by affecting our faculty of representation. Hence, on this view, to talk of the real and its location in space is simply a contradiction in terms. Yet, this was only an anticipatory step toward the view of the latest paper, the German Ideas from 1802, in which he not only regarded matter as the fundamental concept, but even reduced the science of motion to a part of the science of matter: The real in space we call matter and the changes thereof we call movement. Hence, the science of movement, because movement as changing of matter concerns the relations of this, belongs to the science of the relation of matter.45

Thus, Ørsted ended in a position that was as contrary to the one expressed by Kant as one can possibly imagine. Further, to discard Kant’s determination of matter as the movable left Ørsted with a problem. If his doctrine of matter should not be led back to the doctrine of motion, he had to find another fundamental determination of matter on which to ground the doctrine. For this purpose he took recourse to Kant’s transcendental logic in the Critique of Pure Reason, more specifically to the axioms of intuition and the anticipations of perception which form part of the analysis of the possibility of synthetic judgments. Thus, he quoted the axioms of intuition to state that “every object of the external senses must have extensive size” and the anticipations of perception as “in all phenomenal the real, as an object of perception, always has intensive size, i.e. degree.”46 However, Kants axioms of intuition stated that “all intuitions are extensive magnitudes,” not that the object for the external senses must have an extensive size.47 Apparently, Ørsted failed to see the difference between the object for the external senses and the intuition which results when the object has given rise to a sensation and this sensation has been further transformed into an appearance— exactly that mistake which Kant called “the chicanery of a falsely instructed reason, which erroneously thinks of freeing the objects of the senses from the formal condition of our sensibility, and, though they are mere appearances, represents them as objects in themselves.”48 But having introduced ‘the real’, Ørsted continued this argument claiming that “The real in phenomena, then, manifests itself to our senses through the intensive size of its quality, i.e. through a force. Now, matter is the real in material objects…and 44 45 46 47

48

Ibid. p. 82. Ørsted, D. Johann Christian Ørsted’s Ideen…, p. 21. [our translation]. Ørsted, “Fundamentals of the Metaphysics of Nature…,” pp. 59, 60. I. Kant, Critique of Pure Reason [1797], translated and edited by P. Guyer and A. W. Wood (Cambridge,: Cambridge University Press, 1997), p. A: 166. Ibid.

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

109

consequently we must attribute force to matter. No change, and therefore no effect, can happen in sensible nature except through motion, so the force of matter can create nothing but motion, and consequently it must be a motive force.”49 In this way Ørsted made matter a more fundamental concept than motion and force. In his own opinion the argument also did away with the “contingency” of Kant’s matter concept. Toward the end of Dissertation which—as mentioned above—in structure and argument differs substantially from Kant’s Metaphysical Foundations, Ørsted gave a very brief outline (in fact no more than a table of contents) of Kant’s book and then proceeded to give it “a critical scrutiny.”50 What concerned Ørsted was that Kant only wanted to investigate from a priori principles how matter is possible as object of experience, not that matter is possible. The latter, Kant claimed, we only know by empirical observation. In Ørsted’s opinion this claim flawed Kant’s whole endeavor which was to deduce the laws of external experience altogether a priori and thus with apodictic certainty. “Therefore, Kant offends against his own principles when he states that the notion of matter must be derived from experience, for in this way its universality and necessity will be lost. Further, if the very notion of matter is contingent, the deduced laws of nature cannot become necessary.”51

7. A PRIORI LAWS OF CHEMISTRY In Metaphysical Foundations Kant had concluded that as long as no underlying principle of how matter interacts with matter can been found, an a priori deduction of the principles of chemistry will not be possible. All that we know about chemical changes will therefore be based on observation and not on first principles and chemistry will remain “a systematic art, an experimental doctrine, but never proper science.”52 Ørsted touched on this question much later in his career.53 It clearly had his interest in Fundamentals, too, but he offered no solution. Instead he admitted that “To indicate the causes of chemical action with strict philosophical precision, is not so easy as it might seem at first glance now that we have all the fundamental laws of the action of matter.”54 He went on to demonstrate that attempts to solve

49 50 51 52 53

54

Ørsted, “Dissertation on the Structure of Elementary Metaphysics…,” p. 60. Ibid pp. 93–96. Ibid p. 96. Kant, Metaphysische Anfangsgrunde…, p. A: V. For example in First Introduction to General Physics: “The theory of motion has been almost completely transformed into mathematics. The theory of force awaits the inventive mind which can lead it to the same point” Ørsted, “First Introduction to General Physics [1811],” in Jelved, Jackson, and Knudsen, Selected Scientific Writings…, p. 296. Ørsted was more confident in 1846 when he gave a lecture to an assembly of natural scientists in Kiel: “One must admit that the chemical laws of nature are laws of reason [Fornuftslove] like the mechanical laws .…Nowadays we already perceive the basis for the mathematical laws of the connections of material parts to one another, and the relations between shapes and constituents are beginning to dawn on us; I say dawn…because [recent discoveries] apparently only make up the first light for that which we may expect in the future.” Ørsted, “Fundamentals of the Metaphysics of Nature…,” p. 71.

110

K. NIELSEN AND H. ANDERSEN

this problem made previously, by Eschenmayer and Schelling respectively, were miserable failures. In Dissertation, however, Ørsted boldly presented ideas concerning the relationship between cohesion and the two fundamental forces which he thought could be developed into the foundation for chemistry found lacking by Kant. According to Kant, the forces responsible for cohesion were not fundamental, but were derived forces, and as such only to be known to us through empirical investigations. Ørsted immodestly confessed that he originally intended to follow in Kant’s footsteps, but “when I thought it over more carefully I was forced to leave that trail.”55 Ørsted’s new and apparently original idea was, for a given body, to regard cohesion as a result of the combination of the attractive and the repulsive forces taking place within that body, since it was obvious that the repulsive force—in opposition to its attractive counterpart—was not active outside the limits of the body. Chemical action between two bodies, A and B, required close proximity and Ørsted now explained how, at the point of contact, the attractive force of body A will interact (in the manner of the combination of such forces which give rise to cohesion) with the repulsive force of body B. And vice versa. If the new forces of cohesion were strong enough to overcome the original cohesive forces, the bodies, according to this explanation, could enter into a new chemical combination and thus produce a new body. “Here we have only aimed at demonstrating the possibility of an a priori chemistry. However, through a more thorough investigation along these lines, I hope we shall one day be able to know the whole doctrine of chemical affinities.”56

8. NEWTONIAN MECHANICS As has already been stated, a major motive for Kant in writing Metaphysical Foundations was the wish to demonstrate that there is no contradiction between the epistemology laid out in his Critique of Pure Reason and the fundamental laws of mechanics as formulated in Newton’s Principia. In his second chapter, termed “Metaphysical Foundations of Dynamics,” Kant treated matter with relation to the categorial class “quality.” He discussed various forms of forces (friction, cohesion, elasticity, chemical action, surface forces, penetrating forces) and came to the significant conclusion that his own attractive force was identical to Newton’s gravitational force. In the following chapter, termed “Metaphysical foundations of Mechanics” he took great pains to derive Newton’s three laws of motion from first principles. In Ørsted’s interpretations the emphasis was quite different and the Newtonian element significantly reduced. In Fundamentals Ørsted did mention that the general attractive force is the same as gravity, and then went on to briefly derive the laws of motion.57 In Dissertation, however, gravity is only mentioned in passing,

55 56 57

Ørsted, “Dissertation on the Structure…,” p. 84. Ibid p. 93. Ørsted, “Fundamentals…, p. 64, pp. 68–69.

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

111

and Newton’s laws are dealt with in a single sentence: “From this [i.e. an argument about relative motion] it is possible to deduce the Kantian theorem that ‘the action is equal to the reaction’.”58 As if to substantiate his disapproval of the Newtonian description of celestial mechanics, Ørsted in an appendix speculated about the perpetual motion of a heavenly body moving around a central body. Where Kant had pictured a space filled with matter of a density lower than anything which we can try out in experiments, thus keeping open the possibility of an infinitely fine aether, Ørsted now claimed that “…a vacuum can by no means exist.”59 Therefore a body moving through space would always be pushing matter aside, and in order to be eternal its motion would have to be sustained by some external agent. Ørsted briefly considered if the attractive force from the central body could be such an agent, but realized that his argument required a tangential force. He then suggested a most un-Newtonian explanation, “Some sort of matter emanating from the central body hits the moving body in such a way that the impulse is resolved into two forces; the direction of one goes through the centre of the body, and at the same time it becomes a tangent to the curved line along which the body to be moved will go.”60

9. WHAT KIND OF KANTIAN? We are now in a position to evaluate the relation of the young Ørsted’s natural philosophy to some of the central themes in Kant’s philosophy. For a start, we think that it is important to emphasize that Ørsted explicitly stated that on certain points he deviated from Kant and that he wanted to “improve” on him.61 As we have seen, Ørsted came precariously close to equate “an appearance” with “a thing in itself,” and thus presented himself as a naïve realist. Rather amazing, since he claimed all along to be working on the basis of Kant’s epistemology. Apparently the same inclination towards naïve realism guided Ørsted in his main discontent with Metaphysical Foundations, namely his claim that Kant’s investigation of matter was not based on first principles. “The first point which I think deserves censure is that this most astute man did not deduce the laws of external experience altogether a priori but, on the contrary judged that empirical properties ought to be attributed to matter,…”62 A sweeping and blunt criticism when one considers how explicit Kant was in pointing out that his only purpose was to deduce those properties that matter as a sensual possibility must necessarily possess, in case it is to be sensed by us. In his own words, the conditions for working on the basis of transcendental philosophy were clear: “Now to cognize anything a priori is to cognize it from its mere possibility.”63

58 59 60 61 62 63

Ørsted, “Dissertation…, p. 90 [emphasis added]. Kant, Metaphysische Anfangsgrunde…, p. A: 103ff; Ørsted, “Dissertation…, p. 99. Ibid. Ørsted, “Fundamentals…, p. 48; Ørsted, “Dissertation…, p. 80. Ørsted, “Dissertation…, p. 96. Kant, Metaphysische Anfangsgrunde…, p. A:IX.

112

K. NIELSEN AND H. ANDERSEN

Ørsted’s lack of interest in the compatibility between transcendental philosophy and Newtonian physics must be viewed in the light of his insistence that a philosophical foundation for chemistry can be found in spite of Kant’s claim of the opposite. In Fundamentals Ørsted clearly regretted that such a foundation had not been found, and in Dissertation he put forth a daring and somewhat flimsy hypothesis with the effect that—in his own eyes—he proved Kant wrong on this much discussed point. In fact, Ørsted fully embraced only two sets of ideas from Kant. One was Kant’s analysis of the faculty of understanding, in particular the conviction that all judgments must conform to the structural templates given by the twelve categories. In his analyses in the works referred to here, Ørsted closely followed Kant’s method of analyzing a subject by treating it in relation to the four classes of categories, and Ørsted later reflections on the nature of a natural law seemed to be heavily inspired by Kant’s notion of categories.64 The second idea fully assimilated by Ørsted was Kant’s antiatomism, and the idea that the world is most adequately understood and described through the application of the concept of forces. Thus it seems that already in the years 1798–1799 Ørsted reached the conclusion that he wanted to be a chemist and that he wanted to understand not only chemical processes but the entire firmament on a “dynamical” basis. This leaves us with the impression not of a philosopher—or at least of a philosopher of such unsophisticated breed, that finer distinctions concerning his philosophical susceptibility appear a bit moot—but of a determined, gifted, pragmatic, and very ambitious young chemist.

ACKNOWLEDGMENTS We would like to thank Margit Stougaard Christiansen, Anja Skaar Jacobsen, Kenneth Caneva, and Ole Knudsen for valuable comments on earlier versions of this paper and for their patient urging on the project.

BIBLIOGRAPHY Andersen, H. (1989), En videnskabsmand af rang. Adam Wilhelm Hauch, 1755–1838 (An eminent Scientist. Adam Wilhelm Hauch, 1755–1838), Aarhus: Foreningen Videnskabshistorisk Museums Venner. Beck, L. W. (1969), Early German Philosophy: Kant and his Predecessors, Cambridge, MA: Harvard University Press. Billeskov Janssen, F. J. (1987), “Hans Christian Ørsted,” in F. J. Billeskov Jansen, Egill Snorrason, and Chr. Lauritz-Jensen (eds.): Hans Christian Ørsted, Hellerup: IFV-energi i/s. Bærentsen, K and V. G. Jensen (1977), Om den unge Ørsted (On the young Ørsted), in: Theriaca. Samlinger til farmaciens og medicinens historie, Copenhagen: Dansk Farmacihistorisk Selskab 1977, pp. 11–21. Caneva, K. (1997), “Physics and Naturphilosophie: A Reconnaissance,” in Hist. Sci. xxxv: 35–106. Carrier, M. (1990), “Kants Theorie der Materie und ihre Wirkung auf die zeitgenössische Chemie,” Kant Studien 81: 170–210. 64

H. C. Ørsted, Naturlaerens mechaniske Deel (The mechanical part of natural science) (Copenhagen, 1844).

THE INFLUENCE OF KANT’S PHILOSOPHY ON THE YOUNG H. C. ØRSTED

113

Friedman, M. (1992), Kant and the Exact Sciences, Cambridge, MA: Harvard University Press. Hauch, A. W. (1794), Begyndelsesgrunde til Naturlæren (Foundations of Physics), 2. vols., Copenhagen Hauch, A. W. (1799), Begyndelsesgrunde til Naturlæren. Anden og forbedrede Udgave, Anden Del (Foundations of Physics, 2nd revised ed., part 2), Copenhagen. Jacobsen, A. S. (2000), Between Naturphilosophie and Tradition. Hans Christian Ørsted’s Dynamical Chemistry. Aarhus, Ph.D. Thesis, University of Aarhus (unpublished). Kant, I. (1781/1997), Critique of Pure Reason, translated and edited by P. Guyer and A. W. Wood, Cambridge: Cambridge University Press. Kant, I. (1783/1976), Prolegomena zu einer jeden künftigen Metaphysik, die als Wissenschaft wird auftreten können, Hamburg: Felix Meitner. Kant, I. (1786/1977), Die Metaphysische Anfangsgründe der Naturwissenschaft Frankfurt am Main: Suhrkamp Verlag. Landau, A. (1991), Rezensionen zur Kantischen Philosophie 1781–87, Bebra: Albert Landau Verlag. Plaas, P. (1965), Kants Theorie der Naturwissenschaft, Göttingen. [Kant’s Theory of Natural Science, Boston Studies vol. 159 (1994)] Schäfer, L. (1966), Kants Metaphysik der Natur, Berlin: Walter de Gruyter. Shanahan, T. (1989), “Kant, Naturphilosophie, and Ørsted’s discovery of electromagnetism: A reassessment,” Stud. Hist. Phil. Sci. 20, 287–305. Stauffer, R. C. (1953), “Persistent Errors Concerning H. C. Ørsted’s Discovery of Electromagnetism,” Isis 44 (1953), 307–310. Stauffer, R. C. (1957), “Speculation and Experiment in the Background of Ørsted’s Discovery of Electromagntism,” Isis 48, 33–50. Thuborg, A. (1951), Den kantianske periode i dansk filosofi. 1790–1800 (The Kantian era in Danish philosophy, 1790–1800), Copenhagen: Gyldendal. Van der Wyck (1899), “Kant in Holland,” in Kant-Studien 3: 403–414. Van der Wyck (1903), “Kant in Holland,” in Kant-Studien 8: 448–466. Vannérus, A. (1901), “Der Kantianismus in Schweden.” In: Kant-Studien 6: 244–269. Wielma, M. R. (1988), “Die erste niederländische Kant-Rezeption 1786–1850,” in Kant-Studien 79: 450–466. Williams, L. P. (1966), The Origin of Field Theory (New York: Random House). Ørsted, H. C. (1797a), “Hvorledes kan det prosaiske Sprog fordærves ved at komme det poetiske for nær? og hvor ere Grændserne mellem det poetiske og det prosaiske Udtryk?” (How can the prose language be corrupted by coming the poetic too near? and where are the limits between the poetic and the prose expression?), Minerva, May 1797. Ørsted, H. C. (1797b), “Om Modervandets Oprindelse og Nytte,” reprinted in Ørsted (1920) vol. I, pp. 5–13. Translated into: “Response to the Prize Question in Medicine Set by the University of Copenhagen in the Year 1797: On the Origin and Use of the Amniotic Fluid,” in Ørsted (1998), pp. 3–25. Ørsted, H. C. (1798a), Kemiske Breve. Første Brev, Bibliothek for Physik, Medicin og Oekonomie 14: 152–160 reprinted in Ørsted. (1920), vol. III, pp. 3–6. Translated into: “Letters on Chemistry. First Letter,” in Ørsted (1998), pp. 26–28. Ørsted, H. C. (1798b), Breve over Kemien. Andet Brev, om Varmen, Bibliothek for Physik, Medicin og Oekonomie 14: 313–327, reprinted in Ørsted. (1920) vol. III, pp. 7–13. Translated into: “Letters on Chemistry. Second Letter, on Heat,” in Ørsted (1998), pp. 29–34. Ørsted, H. C. (1798c), “Begyndelsesgrunde til Naturlæren. Anden Udgave ved A. W. Hauch. Første Deel, Kiøbenhavn, 1798” (review of Foundations of Physics, 2nd ed., by A. W. Hauch, Part One), in Kjøbenhavns lærde Efterretninger, reprinted Ørsted 1920, vol. III, pp. 27–32. Ørsted, H. C. (1799a), Breve om Chemien. Tredje Brev, Bibliothek for Physik, Medicin og Oekonomie 16: 16–31, reprinted in Ørsted. (1920) vol. III, pp. 14–20. Translated into: “Letters on Chemistry. Third Letter,” in Ørsted (1998), pp. 35–40. Ørsted, H. C. (1799b), Kemiske Breve, Bibliothek for Physik, Medicin og Oekonomie 16: 165–177, reprinted in Ørsted.(1920) vol. III, pp. 21–26. Translated into: “Letters on Chemistry,” Fourth Letter, in Ørsted (1998), pp. 41–45. Ørsted, H. C. (1799c), “Begyndelsegrunde til Naturlæren. Anden og forbedrede Udgave ved A. W. Hauch. Anden Del (review of Foundations of Physics, 2nd ed., by A.W. Hauch, Part Two), Kjøbenhavn, 1799.” in Kjøbenhavns lærde Efterretninger (1799), pp. 801–815. Reprinted in Ørsted 1920, vol. III, pp. 32–45. Ørsted, H. C. (1799d), “slu*tningen af Recensionen over Begyndelsesgrunde til Naturlæren” (end of the review of Foundations of Physics), Kjøbenhavns lærde Efterretninger (1799), pp. 823–827. Reprinted in Ørsted 1920, vol. III, pp. 45–49.

114

K. NIELSEN AND H. ANDERSEN

Ørsted, H. C. (1799e), ‘Grundtrækkene af Naturmetaphysikken. Tildeels efter en nye Plan’, reprinted in Ørsted 1920, vol. I, pp. 34–78. Translated into “Fundamentals of the Metaphysics of Nature. Partly According to a New Plan,” in Ørsted 1998, pp. 46–78. Ørsted, H. C. (1799f), “Dissertatio de forma metaphysices elementaris naturæ externæ,” reprinted in Ørsted 1920, vol. I, pp. 79–105. Translated into “Dissertation on the Structure of the Elementary Metaphysics of External Nature,” in Ørsted 1998, pp. 79–100. Ørsted, H. C. (1802), D. Johann Christian Ørsted’s Ideen zu einer neuen Architektonik der Naturmetaphysik. Herausgegeben von D. W. H. Mendel, Berlin. Ørsted, H. C. (1844), Naturlærens mechaniske Deel (The mechanical part of natural science), Copenhagen, 1847. Ørsted, H. C. (1849–1850), Aanden i Naturen (The Soul in Nature). Reprinted (4th ed.) Copenhagen: Vinten 1978. Ørsted, H. C. (1920), Naturvidenskabelige Skrifter (Scientific Papers), 3 vols. ed. K. Meyer, Copenhagen: Høst og Søn. Ørsted, H. C. (1998), Selected Scientific Works of Hans Christian Ørsted, translated and edited by K. Jelved, A. D. Jackson, and O. Knudsen, Princeton, NJ: Princeton University Press. Ørsted, M. (1870), Breve til og fra Hans Christian Ørsted (Letters to and from Hans Christian Ørsted). 2 vols. Copenhagen: Th. Lind.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC DAN CHARLY CHRISTENSEN

Understanding the concept of force was one of Kant’s primary concerns in his interpretation of Newtonian mechanics. Similarly, it was the major challenge for Ørsted trying to come to grips with Kantian metaphysics and to establish a natural philosophy that was both dynamical and experimental. In his brilliant introduction to Ørsted’s Selected Scientific Writings Andrew Wilson attempts to determine the evolution of Ørsted’s dynamics throughout his career.1 I agree that Ørsted’s progress was slight and that his explanation of his crucial experiment of 1820 appeared less convincing to contemporary natural philosophers than it did to Ørsted himself. It is the aim of this paper to ascertain firstly, whether Ørsted’s dynamics is as embryonic as asserted, and if this is the case, why was it so? To elucidate this question I shall, secondly, dive into his theory of music, which was developed from 1807 to 1839. I shall argue that Ørsted’s concept of force when applied to tones must have a dual character in order to evoke pleasure. According to the Pythagorean tradition tones are physical phenomena that obey the rules of mathematics. In order to arouse an aesthetical experience, on the other hand, they must at the same time evoke ideas of the imagination that defy these constitutive rules. To establish a concept of force that fitted physics as well as aesthetics he needed to reconsider the classical body–soul dichotomy. Hence, his concept of force had to be phenomenal (governed by laws of nature) and noumenal (imaginative and free) at the same time. His suggestion as we shall see was a dualistic notion of tones that embody constitutive and regulative principles simultaneously while moving unconsciously between the phenomenal body and the noumenal soul. For this purpose Ørsted needed a concept of force that made possible the unification of the physical and the aesthetical aspects of nature. 1. DID ØRSTED’S CONCEPT OF FORCE MOVE BEYOND KANTIAN METAPHYSICS? As a student at the faculty of medicine in Copenhagen Ørsted had the option of following two courses on natural philosophy; either the lectures of Ludvig Manthey, professor of chemistry at the independent Academy of Surgery, or those of 1

Selected Scientific Works of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen with an introduction by Andrew D.Wilson (Princeton, NJ, 1998), pp. xv–xl.

115 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 115–133. © 2007 Springer.

116

D. C. CHRISTENSEN

Nicolai Aasheim, professor of experimental physics at the University. Manthey’s lectures followed textbooks by F. A. C. Gren, the German chemist, whereas Aasheim used The Principles of Natural Philosophy by A.W. Hauch, the only Danish physicist at the time of some international reputation.2 Ørsted preferred the former professor, although he was an outsider in two respects. Manthey’s chair did not entitle him to perform as an examiner at the University, and more importantly, Gren’s textbooks, albeit in self-contradictory ways, were impregnated with Kantian dynamics. In contrast, Hauch’s Principles was loyal towards atomism as propagated by the French Academy of Science, recently metamorphosed into the National Institute. Consequently, Ørsted could not help being pushed into the controversy between German metaphysics and French science about dynamism against atomism/ the imponderables. There were two Ørsted brothers, living like inseparable twins and sharing accommodation as well as philosophical leanings. After graduation in pharmacy 1799 Christian, the elder brother, decided to make Kantian metaphysics the bedrock of his scientific career. The crux of the matter was that Kantian philosophy was gaining a foothold in Copenhagen, even if confined to staff and students of divinity and law, among them Anders, the younger brother, a student of moral philosophy. The two Ørsted brothers, who remained very close throughout their lives, had become editors of Philosophisk Repertorium, the leading journal disseminating Kantian philosophy in Denmark. Christian contributed the only, but lengthy and thorough, article on natural philosophy, which was a harsh critique of atomism, and soon elaborated it into a doctoral dissertation. It was an exceptional example of a natural philosophy modelled on Kantian metaphysics. He published thrice on the subject 1. A Review of Hauch’s Principles confronting it with Kantian metaphysics, in Philosophisk Repertorium 1799, pp. 145–224. 2. His dissertation for the doctorate in philosophy (not medicine), Disseratio de forma metaphysices elementaris naturæ externæ, Copenhagen, 1799. 3. A German translation and revision of his dissertation edited by M. H. Mendel, Ideen zu einer neuen Architektonik der Naturmetaphysik, nebst Bemerkungen über einzelne Theile derselben, Berlin, 1802. Christian enthusiastically promoted Kantian epistemology from Kritik der reinen Vernunft as well as from Metaphysische Anfangsgründe der Naturwissenschaft. This means that Ørsted perceived matter as polar forces filling space. While attractive forces work at a distance, the repulsive force works on the surface of adjacent bodies. He mildly criticised Kant for deducing part of his epistemological apparatus from physical phenomena rather than from pure metaphysics. I suppose, that what Ørsted had in mind, was the fact that Kant had deduced his scheme from Newton’s laws of motion, the supreme scientific status of which Kant took for

2

F. A. C. Gren, Grundriß der Naturlehre, vols. 1–2 (Halle: 1797). F. A. C. Gren, Grundriß der Chemie, vol.1 (Halle: 1796). Adam W. Hauch, Begyndelsesgrunde til Naturlæren, 2. forbedrede Udgave, vols. 1,2 (Copenhagen: 1799).

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

117

granted.3 Newton’s laws of motion were recognised as universally true, but everybody including Newton himself was at a loss to explain the cause of gravity. Hence, his laws dealt with the exact phenomenal effects of unknown causes. The phoronomy of celestial bodies was calculable by constitutive rules, i.e. mathematical equations. Gravitation, the attractive force, was external to passive bodies and operative at a distance. When in Opticks Newton transformed the celestial forces into a corpuscular theory of all matter, he introduced a repulsive force acting by immediate contact. By analogy Newton’s corpuscular theory was based on the coexistence of matter and forces, and he repudiated the idea of conceiving forces as occult qualities of matter. Contrary to Newton and the Newtonians, Kant considered matter as that which fills space continuously by means of attractive and repulsive forces acting in space. Empty spaces cannot be intuited and forces cannot act in empty space. So, to Kant matter is not atomistic and porous, but the result of continuous action of polar forces in space. Substance is an object of our intuition, but the reason why we intuit it as such is because “substance” is one of the categories of the understanding. We cannot know that Newtonian corpuscles are substantial, because they defy intuition in time and space. The only “thing” we can intuit is resistance when we try to penetrate “it,” so resisting forces may be all there is. Furthermore, that substance is intuited is even empirically contested by water, which is only substance when frozen. Heat it up and it liquefies and eventually evaporates. So, whereas to the French Newtonians matter (and by inference corpusles/ atoms) has extension and is intuited as quantity in space, forces are intensive and cannot be intuited in space. Conversely, to Ørsted as to Kant—and we could add Faraday4—matter is reducible to the interaction of forces.5 This is the crucial difference between Newtonian mechanics and Kantian/ Ørstedean dynamics. Mechanics operates by means of external forces acting upon inert matter. Gravity is proportional with mass, which is quantitative.6 Dynamics, on the other hand, operate with matter as an empirical phenomenon the metaphysics of which reduces it to forces which qua forces (causes) have no geometrical extension and no numerical value. Since forces cannot be intuited in time and space they belong to the realm of noumena.7 3

4

5

6

7

This is equivalent to Harold Bloom’s advancement of canonical criteria from the text of Shakespeare rather than evaluating literary quality on the basis of criteria constructed by theorists of literature, see Harold Bloom, The Western Canon, The Books and School of the Ages (New York, 1994), p. 40. David Gooding, “Metaphysics versus Measurement: The Conversion and Conservation of Force in Faraday’s Physics,” Annals of Science, 37 (1980), pp. 1–29. Selected Scientific Writings of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen, p. 379, “nothing more than these forces is necessary even for the filling of space.” The fundamental difference between bodies is perfectly explainable by “the degree with which forces fill space, in the predominance of one of the forces, and in its mode of action.”, ibid. p. 383. “We say, “Gravity is everywhere proportional to the quantity of matter.” But what is the quantity of matter? How do we measure it? By weight?—But what is weight other than the result of gravity?—In other words, the phrase “general gravity is everywhere proportional with the quantity of matter” ought to be “gravity is everywhere equal to gravity.” But this is a circle that explains nothing. How, therefore, can we say quantitative gravity?—Everything must be qualitative.” J. W. Ritter, Fragmente aus dem Nachlass eines jungen Physikers, edited by Steffen and Birgit Dietzsch (Hanau/Main: 1984), p. 67. “That the possibility of fundamental forces should be made conceivable is a completely impossible demand; for they are called fundamental forces precisely because they cannot be derived from any other forces, i.e. they cannot be conceived.” I.Kant, Metaphysische Anfangsgründe der Naturwissenschaft, in Kant’s Werke, vol. 4, (Berlin, 1911), p. 513.

118

D. C. CHRISTENSEN

As a consequence, thirdly, Ørsted changed the architectural order of Kant’s Metaphysische Anfangsgründe der Naturwissenschaft by giving dynamics priority above phoronomy and mechanics. This was not only justified by the metaphysical change of priority of moving forces above motion of matter in space. It was made necessary by Ørsted’s focus on chemistry, heat, and light, and above all the forces supposed to be operative in magnetism and galvanism, which latter arrived too late to attract Kant’s attention. Kant’s architecture followed the sequence of his analysis of Newton, whose law of inertia implies that matter is a passive object of polar forces. Gravity, the attractive force, is a product of mass and velocity of bodies operative at a distance, and is indirectly calculable by measuring the latter; the force itself is an unknown thing-in-itself. Consequently, Kant’s architecture puts phoronomy first (matter in motion) and dynamics (the moving force) later. By doing so Kant follows Newton, in so far as his architecture conforms to the intuition of matter as substance, and not to the metaphysics of dynamics. Ørsted—despite his accept of Kant’s analysis—finds that the architecture ought to deal with causes and effects separately and in this logical sequence, and furthermore to undertake a complete analysis of both according to the categories of the understanding. But, alas, the concept of force is not susceptible to analysis by the categories, because it cannot be intuited in time and space. So, Ørsted’s architecture cannot include a chapter on force as a parallel to Kant’s “Dynamics.” Hence Ørsted’s article in Philosophisk Repertorium is subdivided into “Phoronomy”—the doctrine of pure motion—and “Hylology”—the doctrine of pure matter. His doctoral dissertation maintains this division, but reverses the sequence and adds a chapter on the applied doctrine of motion, i.e. mechanics. Mendel’s German edition of Ørsted’s revision of his dissertation also maintains this structure (and the sequence of the dissertation) and furthermore adds an appendix on the ether, which is reminiscent of Kant’s so-called ether deduction in his Transition Project.8 Kant emphasised that natural laws can never be constructed from noumena (noumenorum non datur scientia). Laws of nature are established from phenomena, i.e. by the applying the categories of the understanding on appearances, or on things-for-us, not on things-in-themselves. Phenomena can be subjected to constitutive rules. The law of gravity is known only indirectly by its effect; gravity in itself is an unknowable noumenon and cannot be subjected to constitutive rules. The electrical force is the cause of a shock induced on the human body and the repulsive force is the cause of resistance when we try to penetrate matter. Strictly speaking, there can be no concept of forces since they are not susceptible to the categories of the understanding. We are surrounded by solid bodies, liquids,

8

Kant’s ether deduction, although not published at the time, must somehow have been known by Ørsted and his circle in Berlin during the winter 1801–1802. I find it most likely that Ørsted was made aware of Kant’s Transition Project by Markus Herz, who together with his wife Henriette Herz hosted the so-called Kränzchen circle in Berlin once every week. Ørsted was a regular guest in Herz’s home during the winter 1801–1802. According to his diary (n. 23) Ørsted joined the Kränzchen fourteen times. Markus Herz had been Kant’s student in Königsberg, and being Kant’s preferred expert in philosophy was appointed the examiner of Kant’s dissertation in 1770, Uwe Schultz, Immanuel Kant in Selbsterzeugnissen und Bilddokumenten,(Hamburg, 1965), p. 86.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

119

and vapour as substantial appearances, but from a metaphysical point of view these objects are nothing but attractive and repulsive forces in conflict. Hence, I suggest, Ørsted—like Faraday—in a Kantian perspective was justified in presupposing the existence of forces, while at the same time remaining tacit about the essence of forces. Nevertheless, from 1806 when Ørsted had been appointed professor extraordinarius and had obtained a state grant he planned to publish a comprehensive textbook on natural philosophy. In contrast with the Naturphilosophen it was not his aim to indulge in speculation, but to write a textbook of physics and chemistry from a dynamical point of view. Apparently, he found himself in insurmountable trouble attempting to complete it and his efforts petered out. Only one printed copy survives, probably a first proof.9 His intention, to judge from this, was to divide physics into two sections, one dealing with material bodies being moved by external forces (doctrine of motion, or mechanics), and the other dealing with forces (dynamics). This latter section again was subdivided into two parts: lower dynamics comprising chemistry, and higher dynamics comprising electricity and magnetism—working at a distance—and light and heat—being unbound by matter. He maintained his—and Kant’s—assumption that all forces are reducible to two fundamental forces (causes), but that they are inaccessible by science, since their effects only are empirical. So Ørsted’s Kraftlære (Dynamics) went a step further from his dissertation in so far as he now fully carried out the division between mechanics and dynamics while subdividing the latter into lower and higher dynamics. This outline of natural philosophy, apparently, was also brought out in his lectures to judge from notes taken by one of his students.10 He probably included a large part of this proof and other raw drafts of higher dynamics, if any, in Ansicht which he finished on his second grand tour abroad 1812–1813. The architecture of Ansicht, however, did not conform to the outline of his stranded proof and draft. In fact, he advanced a few hypotheses of dynamism based on a number of experimental results of the effects of forces. He iterated his conviction of the polarity of forces. Firstly, Lichtenberg’s figures, the visual manifestation of discharges from Leyden jars, showed different patterns transmitted by negative and positive electrical forces which seemed to propagate according to the rules of geometry as ramifications on the one hand and as concentric radiation on the other hand.11 Secondly, Ritter’s tortuous experiments with galvanic forces on his private sense organs showed effects of a dual nature in conformity with the negative and positive charges of galvanism.12 A third piece of evidence of the polarity of forces was Ritter’s discovery of ultraviolet light outside the spectrum opposite to the end where Herschel had demonstrated the existence of ultrared light.

9

10 11 12

“Kraftlæren,” no front page, no place and year of publication. 224 pp. The only preserved copy is in private ownership. That Ørsted is indeed the author is proved by his reference (p. 5) to “my” lecture “What is Chemistry?” (1805). Hermansen, Ørsted’s lectures, 1826, Ørsted-Samlingen 98, Royal Library, Copenhagen. Selected Scientific Writings of Hans Christian Ørsted…, p. 381. Ibid. pp. 376–377.

120

D. C. CHRISTENSEN

When declaring his indebtedness apart from Kant Ørsted made reference to Gowin Knight’s point atomism, “which assumes that there are two kinds of atoms, with no extension and with the property that those of the same kind repel each other, and those of a different kind attract each other.” Knight’s book was presented to him by van Marum in Holland while on his way back to Denmark from his humiliation in Paris in September 1803 where he had failed to convince the Institut National of the validity of Ritter’s galvanic experiments. Knight’s conception of force saw it as emanating from a point which had by definition no extension and as a consequence was considered immaterial. To Ørsted as to Knight—and we can add Boscovich—matter was not constructed by atoms but was reducible to forces.13 This did not mean that matter did not exist, of course, it only meant that matter was a purely empirical phenomenon. Forces were perceived like gravity in the sense that it was a necessary apriori concept (causation), only its effects were objects of sensibility. Referring to Gowin Knight, Ørsted mentioned Newton’s nutshell theory of matter by citing Priestley’s comment that all the atoms in the universe could be contained in a nutshell, adding that “the gap between this small extension and its complete elimination is not large.”14 Is Ørsted’s view that dynamical physics must emanate from a double track of metaphysics and experimental physics compatible with the critical philosophy? I find it to be in complete concordance with Kant’s Transition Project, which suggests an intersection model between metaphysics (the transcendental analytic and the categories of understanding) and experiments.15 Kant’s posthumous writings, which Christian hardly knew except, possibly, from hearsay, seem to be reflected in Ørsted’s formulation: “The experimental investigation may first have to perform much preliminary work, for nature’s clues can only cause the mind to produce what lies within it.”16 It goes without saying that this methodology is no warrant of scientific knowledge. Ørsted refused to accept Kant’s discouraging judgement that chemistry could ever acquire the status of a proper science. Kant was very impressed with Lavoisier’s chemical revolution and in his Metaphysik der Sitten he compared him to Archimedes and Newton. What impressed him about this anti-metaphysical chemist was his reduction of a chaos of compounds and elements to a few active principles like oxygen being efficacious in respiration, combustion, and calcination as well as in the generation of acids. Kant saw this progress as an indication that chemistry might eventually become a proper science, because science proper consists in reducing the manifold of sensible phenomena to first principles. Now, we know, that Kant in his Transition Project which aimed at bridging the gap between metaphysics and experimental physics was encouraged by Brunonian 13

14 15 16

Gowin Knight, Attempt to demonstrate that all the Phænomena in Nature may be explained by two simple Principles, Attraction and Repulsion (London, 1749). R. Boscowich, Theoria philosophiae naturalis redacta ad unicam legem virium (Viennæ, 1759). Nos.114 and 97 in Ørsted’s library, cf. Fortegnelse over afdøde Geheimeconferentsraad, Professor ved Kjøbenhavns Universitet Hans Christian Ørsteds efterladte Bogsamling (Copenhagen: 1853) Selected Scientific Writings of Hans Christian Ørsted…, p. 380. Michael Friedman, Kant and the Exact Sciences (Cambridge, MA: 1992), ch. 5. Selected Scientific Writings of Hans Christian Ørsted, (n. 1)…, p. 383.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

121

medicine and by Lavoisier’s discoveries both of which seemed capable of reducing empirical chaos into first principles. Whether Kant’s ideas from his then unpublished writings reached Ørsted we do not know, but obviously he adopted Kant’s idea of reflexive judgement as some sort of intersection between a priori principles and empirical experimentation. There is ample evidence that Ørsted’s scientific career was aiming at a unification of these two tracks. Thus he remained faithful to the Kantian recipe: “Thoughts without content are empty, intuitions without concepts are blind.”17 Ørsted’s optimism of a breakthrough of a modern dynamical physics was based on an improvement of the regulative (heuristic and reflexive) rules of Kant’s metaphysics combined with a range of experimental discoveries made by Volta, Ritter, Winterl, and himself. In Ørsted’s mind Kants famous dictum just quoted could easily be paraphrased into saying “Science without experimentation is empty, experimentation without (regulative) ideas is blind.”18 This implies that establishing laws of nature, the zeal of physicists, must conform to the laws of human reason. Thus, Ørsted adds, laws of nature are like thoughts. The idea of nature comprises reason as well as will (action). “And so we see all of nature as the manifestation of one infinite force and one infinite reason united as the revelation of God.”19 This statement resembles closely one of the many drafts of a title page that Kant wrote for his never-completed work: “The highest standpoint of transcendental philosophy in the system of the two ideas by God, the world, and the subject which connects both objects, the thinking being in the world.”20 Does not this amount to saying that the human mind encompasses as the essence of transcendental philosophy two ideas belonging to different realms: one idea of the world (natural laws or necessity) and one idea of God (the highest good or freedom)? All ideas, including the idea of God, spring from the human mind. This quote is on the unity of thoughts of natural and moral philosophy (Kant’s first and second critiques). Can this unity be extended to thoughts of physics and aesthetics (Kant’s first and third critiques)?

2. AESTHETICS AND THE CONCEPT OF FORCE In the first decade or so of the 19th century Christian—as already mentioned—was the only Kantian natural philosopher in Denmark. There were many academics giving their consent to the critical philosophy or at least being open-minded about it, but they were dedicated to other fields of study such as moral philosophy, theology, or aesthetics. The elitist section of these academics and artists met in the Scandinavian Literary Society of which Christian and his brother had been 17 18

19 20

I. Kant, Kritik der reinen Vernunft, in Kant’s Werke, vol. 4 (Berlin: 1911), p. 48. Thomas E.Wartenberg, “Reason and the practice of science,” in The Cambridge Companion to Kant, edited by Paul Guyer (Cambidge: 1992) p. 243. Selected Scientific Works of Hans Christian Ørsted, (n.1), p. 384. Opus postumum, I.III.4, 21: 34, cited by Paul Guyer, “The Unity of Nature and Freedom: Kant’s Conception of the System of Philosophy,” in The Reception of Kant’s Critical Philosophy, edited by Sally Sedgwick, pp. 19–53 (Cambridge: 2000) This long title continues: “God, the world, and what unites both into a system: the thinking, innate principle of man in the world. Man as a being in the world, self-limited through nature and duty”, p. 19.

122

D. C. CHRISTENSEN

elected members. This society offered a platform from which its members could hope to gain support for their views and carve out a career for themselves. An actual division of labour developed between the two brothers according to their academic fields of interest. Access to scientific instruments and laboratory facilities was scarce, so an obvious path for Christian to pursue empowerment was to demonstrate that critical natural philosophy was a potential asset to moral philosophy, religion, and aesthetics. So he plunged into interdisciplinarity for which critical philosophy was so suitable. Indeed, Kant’s own project was not only to demonstrate the limits of pure reason, but also increasingly to reconcile the necessity of theoretical philosophy with the freedom of practical philosophy. In May 1807 Ørsted submitted an essay to the class of physics of The Royal Danish Society of Sciences and Letters21 on his experiments on acoustic figures.22 Court Steward A.W. Hauch, chairing the class of physics and refereeing the essay, found it very interesting; never had he seen the order of acoustic figures and the relationship between the pure tones and the symmetrical figures so beautifully and so exactly demonstrated as by Ørsted. Contributing towards this favourable appreciation was, no doubt, the fact that Hauch was also heading The Royal Theatre including The Royal Chapel, its orchestra. Ernst Chladni’s book Die Akustik, had been introduced to Christian by Dr. Panzer in Leipzig 5 October, 1802,23 and he kept a copy in his private library. But, as Hauch remarked tongue-in-cheek, as far as he was aware nobody had hitherto noticed as the author now did that electricity is at work in the production of acoustic figures. His report stated that the conclusions of Ørsted’s essay might be considered somewhat hypothetical, but that he found it to be fair, all the same.24 While a select committee of the Royal Danish Society of Sciences and Letters were pondering their judgment on Christian’s essay, he had given a paper in the form of a dialogue on aesthetics in Platonic style entitled “On the Cause of the Pleasure evoked by Tones.”25 It was given on 6 February, 1808 in the Scandinavian Literary Society and subsequently printed in its journal.26 This society was 21 22 23

24

25

26

KDVS [Royal Danish Society of Sciences and Letters], Mødeprotokol [Minutes] , 480/1807. Selected Scientific Works of Hans Christian Ørsted, (n.1), pp. 264–281. H. C.Ørsted, Rejsedagbog. Korte Bemærkninger. [Travel diary. Short Remarks] 1801–1804, 1812–1813, Ørsted-coll.15, Royal Library, Copenhagen. The other members of the class of physics joined Hauch’s judgment and found the essay worthy of being published in the transactions of the society. The question was raised, however, whether the essay entitled him to receive the silver medal as a prize of honour or to be proposed a member of the society. It was suggested that Ørsted had been aiming at a membership rather than the silver medal. It was a sombre moment for the physical class, since four members had recently passed away and four members were unable to attend the fortnightly meetings due to their absence from Copenhagen. After much debate Ørsted was elected a member of the Royal Danish Society of Sciences and Letters in late November 1808, KDVS 543, 612, 621/ 1808. Ørsted’s first action as a member was to suggest a prize essay on the relationship between electricity and magnetism, KDVS 701 and 707/ 1809. However, the solution was only provided by himself—in 1820. H. C. Ørsted, “The Same Principles of Beauty exist in the Objects Submitted to the Eye and to the Ear—An Essay”, The Soul in Nature with Supplementary Contributions, pp. 325–351, translated from the German by Leonora and Joanna B. Horner (London, 1852) (reprint London, 1966). Obviously the translation is a misleading paraphrase of the Danish title. SLSs mødeprotokol [Minutes], 1796–, NKS 768d, Royal Library, Copenhagen. Skandinavisk Museum, 1808.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

123

modelled on the same principles as the Royal Danish Society of Sciences and Letters—with the crucial exception that its mathematical class was replaced by an aesthetic class, while both societies held the three classes of history, philosophy, and physics. They had similarly strict requirements of membership and consequently some overlap of members. They shared premises. Ørsted had been elected a member of the Scandinavian Literary Society immediately after his doctorate and he had a majority of supporters among his fellow members; the physicists among them, however, were a minority. Ørsted had already been dedicated to aesthetics a long time. As a schoolboy he had buried himself in a Danish translation of Batteaux’s aesthetics; and he had been awarded his first gold medal at the University for a prize essay on the relationship between prose and poetry. Christian and Anders arranged literary salons in their homes where Oehlenschläger, Anders’ brother-in-law, and Baggesen recited their poems and plays, before they were staged at The Royal Theatre and printed. In 1806 Christian had begun writing his poem “The Airship” which he extended and improved till its final publication in 1836. Throughout his life he was a dedicated member of “The Harmonic Society,” the leading music club in Copenhagen. His dialogue27—strictly speaking a conversation between five interlocutors—was a clear example of his urge to extend his experiments on acoustics into aesthetics to embrace a comprehensive theory of music. Aesthetical philosophy, however, was not the business of the class of physics. He aimed at convincing his interdisciplinary peers that aesthetics was in need of the natural philosophy of acoustics and vice versa. Indeed, what Ørsted initiated here was a lifelong reflection on the aesthetics of music, poetry, and painting from the point of view of natural philosophy or the classical mind-body problem. Eight years ago Christian had been a precursor of Kant’s theoretical philosophy. Now he aimed at a complete reconsideration of aesthetical discourse, if there had ever been one in his native country, and to suggest an interdisciplinary solution of the mind-body issue. His theory was widely acclaimed as a major step forward in the philosophy of aesthetics.28 The choice to discuss aesthetics by means of a Platonic dialogue was very much in the vein of intellectual discourse. Plato’s dialogues were the primordial model, of course, followed up by Galileo’s on the Copernican system and later pursued by Lessing on Freemasonry and Solger and Schlegel on aesthetics. This genre—as opposed to the monologue or lecture—enabled Christian to merge physics and art

27

28

This was not his first dialogue nor the first paper he gave in the Scandinavian Literary Society. In 1805 he read his “Dialoque on Mysticism” the point of which was that some [read: Romanticists/ Naturphilosophen] are attracted to the mysteries of nature only as long as they remain mysteries, whereas they lose interest once it has been demystified by science. True lovers of knowledge on the other hand are attracted by the challenge to overcome mysticism by establishing laws of nature, and this insight does not diminish their enchantment of nature. Paul Moos, Philosophie der Musik von Kant bis Eduard von Hartmann, reprint (Hildesheim/ New York, 1975), pp. 88–93 (rel. to Schelling), 151–152 (rel. to Hegel), 370–371 (rel. to Engel), 533 (rel. to Helmholtz). Eduard Hanslick, The Beautiful in Music. A Contribution to the Revisal of Musical Aesthetics, translated from the 7th ed. (Leipzig, 1885), New York, 1974, passim. Encyclopedia of Aesthetics, edited by Michael Kelly (Oxford ,1998), p. 319.

124

D. C. CHRISTENSEN

and to juxtapose conflicting aesthetical ideas while staging another character as a mouthpiece of his own. The interlocutors represent four opposing views on aesthetics, all of which Ørsted finds himself incited to rectify. 1. Felix is infatuated with music a la Rousseau, but as to the cause of his enchantment he can only echo the standard answer: je ne sais quoi! Music is beyond rational explanation; it is the creation of a genius following no rules. Music inspires the mind to celebrate high ideas and is entirely noumenal. 2. Herman finds that music is generated by laws of reason, which apply to nature and humans alike. He is not representing a particular view on music, but takes it that music mirrors reason. 3. Valdemar sees music merely as a sensual stimulation a De la Mettrie, i.e. as a mechanical intoxication of the body. Composition and reception of music must obey laws of nature. Taste is entirely subjective and haphazard. Music is phenomenal. 4. Julius shares this mechanical view of music holding that the reception of music is sensual, subject to individual tastes, and often illusory as a placebo effect. To settle the issue Alfred (alias Ørsted) introduces the conclusions of his essay on Chladni’s acoustical figures. Who was Chladni? And what were his acoustic figures? E. F. F.Chladni’s (1756–1827) experiments aimed at a visual representation of the figures, not of individual tones, but of sounds. He produced these figures by stroking a plate of glass or metal by the bow of a violin. The plate was supported at its centre and strewn with a thin layer of sand. Acoustic figures would emerge almost instantaneously by the stroke in that the vibrations of the plate would move the sand to form crispations on the tranquil parts (nodal points) of the plate, i.e. the parts kept at ease by the touch of fingers. Chladni was amazed by the regular geometrical patterns of these sound figures. When he divided a quadrangular plate into squares parallel to its sides he could measure the distances between the sides of the plate and each nodal point of the figure. These distances would be analogous to the strings of a piano. Acoustic figures revealed a correspondence between a harmonic sound and a symmetrical figure, or a chaotic relationship between disharmonic sounds and irregular figures, which was therefore easy to explain. Chladni saw his experiments within the framework of mechanics: motions, bodies, vibrating forces. His book was replete with plates and measurements of his figures. Renowned mathematicians such as d’Alembert and Euler made efforts to deduce the natural laws of acoustics on the basis of Chladni’s figures, but to no avail. Ørsted carried out more than a hundred experiments on acoustic figures from the autumn of 1804. In fact, he kept entertaining parties by reproducing acoustic figures throughout his life, and he won quite a reputation as an entertainer in visual music. For its peace, harmony, and joy, Søren Kierkegaard found Christian’s face reminiscent of “an acoustic figure well bowed by Nature.” While Chladni drew his patterns, Ørsted copied his by taking advantage of Lichtenberg’s method, i.e. firstly, pressing a piece of gum-sheeted paper lightly

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

125

on the lycopodium pattern on the electrophor/metal plate, and, secondly, placing very carefully the locypodium on a piece of glass.29 By this method Ørsted’s figures revealed the hyperbolic and asymptotic forms more clearly.30 His copper plates made it easy for everyone to admire the symmetrical figures and to learn that the more perfect they looked the more harmonious were the sounds that had produced them. Right from the beginning Ørsted associated these figures with C. G. Lichtenberg’s patterns and with the galvanic self-torture of the ear and other senses to which Ritter, his closest friend, had daringly submitted himself.31 While Chladni thought that he had explained his figures exhaustively as mechanical motion, Ørsted led by Lichtenberg and Ritter interpreted them in the context of Kantian dynamics. Ørsted, as Hauch had subacidly hinted in his review, found that forces of electricity were responsible for the undulatory making of the nodal lines. In order to prove this he applied fine sem*n lycopodium instead of the coarser sand used by Chladni. Secondly, how would electrical forces be generated at all? The plates whether made of glass or metal were not electrophori. By analogy to Lichtenberg’s experiments—and the two patterns seemed to be related—Ørsted believed that the oscillations produced by the bow were accompanied by weak frictional electricities (+ and −). These would be distributed unevenly on the plate, “undoubtedly negative at the nodal lines, but positive at the dust lines, because the negative lycopodium is attracted to the latter.”32 He was convinced that this was the case because when he tried to knock off the lycopodium “the quiescent lines will be completely devoid of dust, whereas the dust lines proper will retain most of theirs.” Hence a mechanical explanation was inadequate. What spurred his uncertainty, however, was his inability to detect any reaction by means of Coulomb’s electrometer. Ørsted also observed that the acoustic figures varied according to the sort of dust strewn on the plate. The coarse sand would quickly move to the nodal lines, whereas the fine lycopodium was slow and continuous as it moved towards the final dust lines of the figure. Ørsted talked about primary and secondary motions.33 Even if it was difficult to observe the secondary motions due to velocity Ørsted believed to have observed them as being undulations moving quickly both vertically and horizontally across the plate, which was yet another indication that mechanics could not account for the formation of the figures. Ørsted believed that sound, light, heat, electricity and magnetism (the phenomena scrutinised by dynamics) were transmitted by way of undulations. He inferred that acoustic figures owed their formation to electricity, and that their study therefore belonged to higher dynamics. 29

30 31

32

33

G. C. Lichtenberg, Über eine neue Methode, die Natur und die Bewegung der elektrischen Materie zu erforschen (Göttingen, 1777). Selected Scientific Works of H.C. Ørsted, (n.1), p. 277. J. W. Ritter, Beyträge zur nähheren Kenntnis des Galvanismus und der Resultate siner Untersuchung, vols. 1,2 ( Jena 1800–1805), vol. 2, p. 257. Selected Scientific Works of H.C. Ørsted, No.21 “On the Manner in which Electricity is Transmitted (A Fragment)”, p. 214, and No. 28 “Experiments on Acoustic Figures,” pp. 277–278 Ibid. No. 16. “A Letter from Dr. Ørsted of Copenhagen to Mr. J. W. Ritter of Jena, Concerning Chladni’s Acoustic Figures in an Electrical Context,” p. 180

126

D. C. CHRISTENSEN

He wrote to Ritter on the subject, because at the time Ritter was performing a series of galvanic experiments on the senses of his own body. Ritter had noticed that the ear responds differently to the positive and negative conductors of the Volta pile—and receives a shock when they are combined. Ritter responded copiously.34 He assured his friend that the downward-bending waves were charged negatively on the top, positively at the bottom, while the upward-bending ones were charged inversely. At ease these polar charges neutralise each other, but when oscillations start the harmony is disturbed and electrical effects are produced. Ritter was sure that waves of sound propagated like waves of light. In fact, he analogised an octave of tones with a spectrum of light. Furthermore, he claimed that the ear and eye responded to waves due to the curvatures of these sense organs. How human sense organs actually functioned had hardly ever been examined, and their reactions to galvanic charges were definitely a new field of physics. Both friends published their groundbreaking experiments and hypotheses in German journals hoping to conquer more territory for dynamical natural philosophy.35 The electrical force itself, of course, was not observable. Forces are immaterial. Only the effects of forces can be observed. One possible effect was the motion of materials. But if a force caused this motion it would be undulatory, as opposed to mechanical motion. So, Ørsted’s experiments on acoustic figures must be seen in a larger perspective. They aimed at discovering the laws of dynamics and to support Kantian metaphysics, i.e. that only matter is phenomenal, whereas forces are noumenal. On this foundation Ørsted was working out an entirely new structure of natural philosophy as already shown above. Joining Ritter’s vein of research, Ørsted was convinced that a future auditory theory would affirm that sound waves transmit weak electricity which is perceived by the ear and passed on—still by electricity—to the brain by means of nerves acting as conductors of electricity. According to level of tone and sound the mind will feel sorrow, cheer, etc. What profundity unknown to the listener is not hidden in a single chord, what infinite arithmetic in a whole symphony! And now, joined with this, the invisible forms, which appear before our soul in obscure intimations while the notes, flow into the ear. In truth, we can repeat with joy and triumph at the nobility of our spiritual being that what fascinates and enraptures us in the art of music and makes us forget everything while our soul soars on the flow of notes is not the mechanical stimulation of tense nerves. It is the deep, infinite, incomprehensible Reason of Nature, which speaks to us through the flow of notes.36

So, how does Ørsted conclude his theory of music in the united context of aesthetical and natural philosophy? Firstly, Ørsted replaces the dichotomy between material (phenomenal) and spiritual (noumenal) interpretations of music by a dualism. The sound of music affects the senses as well as reason. Instruments 34

35

36

J. W. Ritter to H. C. Ø. 20 May 1803: “Aller Sinnenempfindung liegt Oscillation zum Grunde,” Correspondance de H.C.Ørsted avec divers savants, edited by M.C. Harding (Copenhagen, 1920), vol. 2, p. 40 and pp. 90–91. Kirstine Meyer comments (H.C. Ørsted, Scientific Papers (Copenhagen, 1920), vol.1, p. xlii): “…it cannot be denied that the arguments employed in various places, especially to explain the generation of electricity, bear the stamp of being adapted so as to agree with a previously given result and of building not so much on mathematically or experimentally grounded facts as on hypotheses intuitively advanced. To French and English scientists, in particular, who were not infected with the phraseology of the German school of Nature Philosophy, the form must have been distasteful.” Selected Scientific Works of H.C. Ø , No. 28. “Experiment on Acoustic Figures,” pp. 280–281.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

127

of music obey the laws of nature. Hence sounds are transmitted by air as waves, i.e. phenomenal objects of nature. This implies that composer, musician, and listener create, perform, and perceive these phenomena objectively. However, none of them are conscious of the mathematical character of sound. So even though melodies, harmonies, crescendos, and tempo, etc. can be perfectly described by constitutive rules, composers, performers, and music-lovers are completely unaware of the mathematics of acoustics. What these people have in mind are ideas of reason, which belong to the noumenal realm. Ideas of reason are inseparable from the sound of music, so phenomena and noumena accompany one another. Sound is transmitted from the sense organ via nerves to the mind. But in the dialogue Christian (alias Alfred) is silent on the possible part of electricity. On the cognitive level tones are physical, necessarily governed by constitutive rules, whereas on the aesthetical level they provide ideas conducive to free imagination and reflection. The linkage between the two levels is established unconsciously. Secondly, he takes advantage of the concordance of the Chladni figures between visual and audible beauty. Figures and waves embody objectivity, and human reason prefers geometrical order to chaos. The experience of beauty is therefore neither subjective nor illusory, but founded on laws of nature. Ørsted goes on to explain that enchantment of music depends on the satisfaction of our senses and the stimulation of our imagination. Imagination is the human capacity to make sense impressions provide ideas of reason for pleasure and reflection. Alfred makes his point visual by suggesting that humans admire symmetrical forms like the circle and well-proportioned figures while they abhor squiggle and chaos. Unconsciously, so runs Ørsted’s argument, echoing neo-Platonism, we tend to imagine the “hidden” ideal form when we are faced with the appearance of an imperfect one whether in nature or in art. This we call taste, and Rousseau was at a loss to explain what it is. He just shrugged his shoulders admitting je ne sais quoi. Now Ørsted suggests a theory of music, which integrates the two realms of science and art. Tones and intervals are phenomenal and susceptible to mathematics. But mathematics is neither what Mozart creates nor what we listen to. If Mozart had had to apply the mathematics of tones, time would not have allowed him to compose even a single symphony. And yet music is replete with mathematics. The phenomenal aspect of music is necessary but irrelevant for composer and concertgoers alike. What the minds of Mozart and the music-lover share are the noumenal, the ideas of beauty and reason. The noumenal aspect of music is free—and conscious. Ørsted’s theory of music bridges the gap between necessity and freedom, between science and aesthetics, between the phenomenal and the noumenal. The two realms coexist inseparably and unconsciously, since aesthetics are conditioned by science, but not only by science. So, in this respect Ørsted saw no difference between himself, the dynamist, and the artist. Together with Ritter he saw himself as a servant of the muses, not as a skivvey of mechanics. As Ritter wrote to Ørsted, “Don’t forget that we are artists!”37

37

Ritter to Ørsted 16–17 August 1805: Vergiss nicht, daß wir Künstler seyn sollen. Kunst aber brauche ich dir nicht zu definiren., in Correspondance…., (n. 34)vol. 2, p. 114; cf. J. W. Ritter, “Anhang,” circa March 1809, in J.W.Ritter, Fragmente aus dem Nachlasse eines jungen Physikers. Ein Taschenbuch für Freunde der Natur, edited by Steffen and Birgit Dietzsch, (Hanau/Main, 1984), pp. 267–286.

128

D. C. CHRISTENSEN

Finally, let me return to the dialogue’s four judges of taste. How has Alfred’s dualistic approach to their topic affected his interlocutors? 1. Felix now understands that music is an unconscious unification of laws of nature and aesthetical imagination, i.e. phenomenal as well as noumenal. He accepts that explaining what to him (and Rousseau) was a mystery by laws of nature does not deprive art of its enchantment. Quite the opposite. It amplifies it. 2. Herman, too, the idealist, is persuaded by the dualistic theory of music and gives up his idea that music is exclusively a mirror of reason. He now appreciates why music unconsciously links body and mind. 3. Valdemar, the sensualist, learns that music is not merely a stimulation of the senses, but that sensual experience is inextricably intertwined with the imagination of ideas of reason. He realises why music constitutes an unconscious link between necessity and freedom. 4. Julius has to recognise that individual taste is anchored in laws of nature, and watching the acoustic figures convinces him that beauty is not illusory, but concords with the visual representation of tones. In 1839 Ørsted wrote two more dialogues on the same topic.38 By that time he had had the opportunity to follow up the subject. His third grand tour of Europe which, upon his discovery of electromagnetism took the form of a triumphal procession brought him into personal contact with Charles Wheatstone, John F. W. Herschel and Michael Faraday who shared his interest in acoustic figures.39 We know from a letter to Wheatstone, that Ørsted had demonstrated the experiments on acoustic figures in Paris, but at that stage, apparently, he had waived his former tentative electrical explanation of the secondary vibrations of the crispations, which he had never mentioned in his dialogues. The crucial problem, however, remained unsolved and attracted the attention of Wheatstone who failed to solve it. The problem, as you will remember, was this: the primary motions of dust towards the formation of heaps and the secondary, internal vibrations in the now stationary heaps cannot be caused by the same force. Uncertain that the secondary motions are effects of frictional electricity, Ørsted now (1823) inferred “that these minute vibrations are communicated by the sounding body to the air, and that they are the cause of the timbre of the tones.”40 The problem puzzled Faraday and during six months in 1831 he carried out a series of experiments at the Royal Institution. Like Ørsted, Faraday had a hunch that a mechanical explanation of the phenomenon had to be ruled out as 38 39

40

See below. Ørsted kept the publications of these natural philosophers in his private library and exchanged letters with them. Fortegnelse over afdøde Geheimeconferentsraad, Professor ved Kjøbenhavns Universitet Hans Christian Ørsteds efterladte Bogsamling, Copenhagen 1853, especially pp. 17–19. The following authors relevant to aesthetics, music and sound are represented in Ørsted’s private library (containing according to this catalogue 6,500 volumes + an unknown number inherited by the family): The Encyclopédie, Chladni, Euler, Gall, Lichtenberg, d’Alembert, Diderot, Rousseau, Kant, Fichte, Schelling, Ritter, Fr. Schlegel, Plato, Kierkegaard, Burke, Faraday, Herschel, Wheatstone. Correspondance…, (n. 34), vol. 2, pp. 320–328, 385–405, 588–597. Ørsted to Wheatstone 12 June 23, Correspondance…, (n. 34), vol. 2, p. 590.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

129

contradictory. Instead, like Ørsted, he assumed that the secondary vibrations in the heaps were caused by waves of air generated by the primary motions. This effect he termed acoustic induction. Unlike Ørsted, Faraday proved this to be the case by repeating the experiment in a vacuum. Result: no vibrations!41 At this stage, Faraday had familiarised himself with Ørsted’s 1806-essay on the visualisation of sound.42 Faraday seems to have been as hopeful as his Danish colleague about the groundbreaking consequences of this new insight.43 The second dialogue, “The Physical Effects of Tones,”44 was staged as a reunion between the interlocutors of the 1808-dialogue pondering questions about the nature of music from a scientific point of view. Now as then Alfred is the knowledgeable mouthpiece of Ørsted. At the end of the evening Valdemar chides Alfred for having changed his mind. As far as he remembered Alfred used to be a spiritualist cherishing the human imagination as the gateway to beauty and higher ideas. Now he saw in him a materialist only. Rejecting this Alfred claims that now as then he merged the two otherwise opposed views. The friends at the reunion held great expectations as to Ørsted’s mature insight into acoustics. However, he modestly confesses that he has little to add to his first theory of music. Science does not yet know how these transmissions from ear to brain to muscle take place. However, he finds good reason to think that dynamical forces are active here, such as small changes of temperature, electricity, and magnetism. Whether the soul is connected to the body by the nervous system, Alfred just cannot tell, but in the end they agree that different kinds of passions in the soul affect the well-being of the body each in their own way.45 The discussion then peters out and the dialogue metamorphoses from the expected discourse on aesthetics into a thorough exposition of Ørsted’s views on the metaphysics of natural science. Having unfolded his dualistic theory of music in the first dialogue and updated it in the second one, Ørsted probably found that he had exhausted his mind in that particular field. 41

42

43

44

45

Ryan D. Tweney, “Stopping Time: Faraday and the Scientific Creation of Perceptual Order,” Physis, vol. 29 (1992), pp.149–164. Selected Scientific Works of H.C. Ø., No. 21, translated into French, “Sur la propagation de l’electricité, Journal de Physique,” 62 (1806) and into English, Nicholson’s Journal, 15 (1806), cf. L.Pearce Williams, Michael Faraday, New York 1965, p. 179. Some six months after his discovery of electromagnetic induction, Faraday had a sealed note deposited in the Royal Society, stating, among others, I think also, that I see reason for supposing that electric induction (of tension) is also performed in a similar progressive time. I am inclined to compare the diffusion of magnetic forces….to the vibrations upon the surface of disturbed water, or those of air in the phenomena of sound; i.e. I am inclined to think the vibratory theory will apply to these phenomena, as it does to sound and most probably to light.- By analogy I think it may possibly apply to the phenomena of induction of electricity of tension also. – These views I wish to work out experimentally: but as much of my time is engaged in the duties of my office, and as the experiments will therefore be prolonged, and may in their course be subject to the observation of others; I wish, by depositing this paper in the care of the Royal Society, to take possession as it were of a certain date, and a lone right, if they are confirmed by experiments, to claim credit for the views at that date: at which time as far as I know no one is conscious of or can claim them but myself, ibid. p. 181. H.C. Ørsted, The Soul in Nature, pp. 352–371. Rather than the translation from the German edition to the English, “The Spiritual in the Material,” I would suggest “Mind and Matter.” Ibid. p. 370.

130

D. C. CHRISTENSEN 3. ØRSTED’S AMBITION TO AMALGAMATE NATURAL AND AESTHETICAL PHILOSOPHY

The third dialogue on mind and matter takes place in the same house the following evening—we are still in 1839—and takes an unexpected turn. Ørsted states in an editorial note that this third dialogue was written as a continuation of the second, but having finished writing it he realised that it had a broader philosophical aim.46 Notwithstanding, he must have found his conclusive remarks so noteworthy as to deserve to be placed first in his The Soul in Nature. The obvious question then is: What made this third dialogue so important as to justify it as the introduction to his philosophical testament? Ørsted commences with a popular adaptation of the first principles of dynamical theory based on Kant’s Metaphysische Anfangsgründe der Naturwissenschaft. Using easy metaphors and simple examples he makes his interlocutors understand that things are not what they immediately seem to be to our senses and that—on reflection—matter is nothing but space filled by polar forces resisting attempts to penetrate its space. Hence what we think is a material world should be understood as the effects of forces which constantly change the form of the objects of perception. All nature is in a constant state of flux. Hence our senses deceive us because they only perceive appearances of the outer world. To get to grips with this outer world, our minds make unconscious use of the apriori categories of the understanding, our capacity to make judgments from experience, and that the causes of dynamics cannot be experienced by the senses, only the effects of forces, etc. For our comfort, fortunately, the laws of nature governing these current changes are themselves universal and unchangeable. Ørsted concludes this first part by quoting Schiller: What the Spirit promises Nature performs!47 Secondly, he confronts the alternative idealistic positions of Fichte and Schelling, now being represented by Herman and Felix respectively, although the two German philosophers are not mentioned by name. The Schiller-quote arouses the immediate suspicion of the interlocutors who have now been assigned new roles. Aren’t we then just constructing nature in accordance with our own preconceived ideas, so that the laws of nature and the phenomena of nature are mere tautologies? This was the common charge against Fichtean idealism, now represented by Herman. Against his fierce opposition Alfred, still Ørsted’s mouthpiece, turns it back on him to exemplify that the outer world does indeed offer independent resistance, 46

47

Ibid. pp.1–27. The English translation of The Soul in Nature, unfortunately, does not provide the correct information on the years these dialogues were written. Furthermore, it adds to the confusion that the first (1808), pp. 325–351, and the second (1839), pp. 352–371 dialogues are printed together under a mutual title. Hence the opening section of the third dialogue (1839), pp. 1–27, which Ørsted himself found so fundamental that he decided to make it introduce the entire collection, appears to be meaningless. Secondly, the translation itself is flawed. It was made not from the Danish original, but from the German translation, since the Horners did not know Danish. To give but one grave example from “The Principle of Beauty perceived by the Eye and Ear”: The English “third” is used as a translation of both the Danish “terts” and “treklang.” Hence this sentence: “The most beautiful of all harmonies, in itself, is the chord of the major third.” is nonsensical.“Third” must be replaced by “triad.” Unfortunately, this and other errors pop up too often. Thanks to Miss Nina Bundgaard, a student of mine, for making me aware of this. Ibid. p. 13: Was der Geist verspricht leistet die Natur.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

131

since their disagreement is genuine, not a construction by one of them. In response to Herman’s constructivist idea Alfred declares himself a Kantian dualist distinguishing between human reason and law-bound nature seemingly impregnated with reason. He suggests a double argument: If nature itself did not abide with the laws of nature we could not force them upon it. And if the laws of nature were not part of our reason we would not be able to grasp them. Consequently, humans themselves must be understood as members of the natural world. This statement provokes Felix, now an ardent defender of Schelling’s absolute idealism. To him reducing humans to creatures of the natural world is offensive. Any understanding of man and nature must take its point of departure in the idea of the absolute, i.e. God or the deity. Philosophical systems based either on the necessary regularity of real nature or on ideal freedom of the human will are bound to err. Real nature will be distorted if judged by the standards of the ideal, and vice versa. Hence, to avoid the deadlock of this pair of contradictions Schelling constructs his monist platform of absolute, transcendental idealism. Schelling suggested that his system was adequately communicated by works of art being a synthesis of the real and the ideal worlds. Ørsted’s dualism, according to Felix, included humans in the domain of nature and deprived them of their will power, which was bound to sacrifice human consciousness and freedom. Alfred retorts Felix’ eloquent echo of Schelling by rejecting his absolute platform as the figment of an overwrought brain. Humans have no other choice to get to grips with the outer world than by dynamical natural philosophy, which—as the history of science amply demonstrates—is still according to Ørsted, regrettably, an incomplete and sometimes erroneous adventure. In spite of Schiller’s dictum What the Spirit promises Nature performs, Nature frequently proves science wrong. Naturphilosophen, however, have never put their fantastic speculations to the test, because they have never been seriously devoted to science. Forty years ago Schelling had worked out his Naturphilosophie partly strutting in borrowed plumes. The problem with his system is that nature had no independent position from where it feeds back on transcendental reason. Like astrology it is inclined to consider man the centre of everything. The heavens must turn round him alone; the stars must foretell his fate; for him the whole is created. Do you believe that man would have relinquished these ideas without a study of nature?48 In the end Felix surrenders. It is now obvious that the reason why Ørsted puts so much emphasis on this third dialogue from 1839 as to make it introduce The Soul in Nature is because it epitomises a cardinal point of his natural philosophy. He clearly dissociates himself from Fichte and Schelling and turns to the protagonist of his dissertation: Immanuel Kant. He does so explicitly by rephrasing in a popular style what he sees as the essential reasoning of Metaphysische Anfangsgründe der Naturwissenschaft. But Ørsted goes beyond Kant in so far as he wishes to incorporate his comprehensive theory of music into physics. To accomplish this he takes advantage of Kant’s critique of the mechanical concept of force which ended in a cul-de-sac in so far as a dynamical concept of force defied constitutive rules. It seems to me that Ørsted tried to solve this problem along the dual lines pondered by Kant in his Transition 48

Ibid. p. 19.

132

D. C. CHRISTENSEN

Project, i.e. by pursuing the intersection model between metaphysics and experiment. His experiments on acoustic figures testified to the objective beauty (symmetry) of the effects of tones. There was complete correspondence between the audible and the visual manifestations of acoustic figures. As to the forces behind the tones he had to acquiesce by adopting Kant’s judgment: forces in themselves are unrecognisable and belong to the noumenal realm. So does the human mind or soul, which is susceptible to the aesthetical ideas embedded in the tones. This unification of physics and aesthetics has significant consequences. One of them is the way Ørsted perceived the concept of genius differently from the way Kant had done. In his Kritik der Urtheilskraft Kant had paved the way for Ørsted’s solution to the question “What is the reason of the pleasure invoked by tones?” but Kant had arrived at it from a different approach. Pondering the claims of taste Kant had set up the following antinomy: 1. Thesis: The judgement of taste is not grounded upon concepts; for otherwise it would admit of dispute (decision by means of proofs). 2. Antithesis: The judgement of taste is grounded on concepts; for otherwise, notwithstanding its diversity, the judgement of taste would not even admit contention (making a claim to the necessary agreement of others with this judgment).49 Kant demanded this antinomy to be lifted. He did so by suggesting that “concepts” be used equivocally in the two sentences. “Concepts” in the thesis is taken to mean determinate concepts, i.e. concepts that refer to matters of cognition like the categories of the understanding. The antithesis, however, deals with indeterminate concepts, which refer to the noumenal realm (der bloße reine Vernunftbegriff von dem Übersinnlichen, “the pure concept of reason relating to the supersensible”).50 The latter, of course, are unsuitable to prove whether a cognitive judgement is true or false. They are aesthetical ideas conveying pleasure and beauty in the mind of the art lover. Without them it would be impossible to establish any discourse on aesthetical judgement, whether agreement is attained or not. It should be noted that from a physical point of view tones can be judged as phenomena by determinate concepts and obtain exact mathematical description. But this is not the issue at stake now. We are only concerned with the aesthetical judgement of tones (melodies and harmonies), which do not represent objects that can be intuited in time and space, albeit do invoke ideas of the imagination. So, in conclusion aesthetical judgement is possible by application of indeterminate concepts, which pertain to the noumenal character of the work of art. In extension of this solution Kant infers that whilst determinate concepts like the categories of the understanding can be systematised in a set of rules, indeterminate concepts cannot. This is the reason, according to Kant, why the creation of art as well as the enjoyment of art is unaccountable for in terms of a given set of rules. Art is a manifestation of a genius, whose talent is the gift of nature. There are no rules to be learnt. The “rules” of art are imposed upon the artist behind his back as it were. Genius is characterised by Kant in the following ways: firstly, originality 49 50

I. Kant, “Kritik der Urteilskraft” §56, Kants Werke, vol.v ( Berlin, 1913), pp. 338–339. Cf. Paul Guyer’s discussion of “a supersensible substratum” in his Kant and the Claims of Taste, (Cambridge (MA.)/London, 1979), pp. 337–345.

ØRSTED’S CONCEPT OF FORCE AND THEORY OF MUSIC

133

is a must. Secondly, its work of art is exemplary and thus prone to attract plagiaries. Thirdly, it is unable to ascertain any ordained rules governing its creativity; and fourthly, genius cannot be ascribed to scientists because they follow rules that can be learnt, e.g. weighing, measuring, and solving equations.51 Ørsted’s response as to the pleasure of music is very close to Kantian aesthetics. By approaching his problem from two angles, the physical and the aesthetical, he sees tones as acoustic as well as visual (Chladni figures) phenomena that are amenable to a cognitive description by means of determinate concepts. At the same time he locates tones in the noumenal realm, so that an aesthetical judgement must avail itself of indeterminate concepts. The two levels merge unconsciously in the experience of the music lover. Neither the composer nor the concertgoers pay any attention to the mathematical properties of tones on the strictly cognitive (Pythagorean) level.52 Both are carried away by the tones and absorbed in the ideas of the imagination that are hidden in them. However, Ørsted did not agree with Kant’s judgment that a scientist constructing laws of nature was merely following ordained rules. If so, Ørsted would have excluded himself from the Parnassus of geniuses. Although Kant was a true admirer of Newton, he had deprived him of his otherwise recognised status of genius due to his understanding that Newton’s laws of motion had been derived from Kepler plus Galileo plus mathematics. Newton had “just” followed the rules of mechanics and mathematics. Ørsted was not aware of having followed any ordained rules when he submitted his physical and aesthetical papers to the scientific societies. Nor could he tell which rules had been applied when he discovered electro-magnetism.53 And in his debate with Ampère he kept rejecting ideas of applying mathematics to his discovery. To Ørsted the latter’s electro-magnetism was like the tones of music, i.e. forces hidden in the noumenal realm. Not only did Ørsted see himself (and Newton among others) as a genius. He also saw himself (and Ritter, his close friend54) as an artist-scientist. Science he claimed set out from the finite phenomenon and ended up in infinite reason. Art set out from the indeterminate noumenon and ended up in concrete sensual works of art. Science and art to Ørsted were amiable siblings.55 University of Roskilde 51 52 53

54

55

I. Kant, “Kritik der Urteilskraft,” §46, Kants Werke, vol.v, (Berlin, 1913), pp. 307–308. At this point, too, Ørsted is in full agreement with Kant, cf. Kritik der Urteilskraft, ibid. p. 329. To Ørsted it was self-evident that nobody could be taught to make a great discovery. But as a dynamist and a rather bad mathematician Ørsted, the discoverer, could not avail himself of the rules of mechanics and did not apply mathematics. To Ørsted genius meant exceptional independence of mind and he applied the term genius on both great scientists and great artists, since he stresses the similarities between the two branches; cf. H. C. Ørsted, “Nogle Bidrag til at belyse Poesiens og Kritikens Eensidigheder og Overdrivelser i den senere Tid” (“Some Contributions to enlighten the recent Onesidedness and Exaggerations of Poetry and Criticism”), in H. C. Ørsted, Samlede og Efterladte Skrifter, vol. 9, (Copenhagen, 1852), pp. 60–65. Cf. J. W. Ritter’s speech “Die Physik als Kunst” (“Physics as Art”), in J. W. Ritter, Fragmente aus dem Nachlasse eines jungen Physikers. Ein Taschenbuch für Freunde der Natur, ed. by Steffen and Birgit Dietzsch, (Hanau/Main, 1984), pp. 288–320, and Dan Ch.Christensen, “Physics as a Branch of Art—The Romantics in Jena,” in Intersections. Art and Science in the Golden Age, edited by Mogens Bencard (Copenhagen, 2000), pp. 18–31. “Science and art on their highest peak approach one another amiably as siblings,” in H. C. Ørsted, “Over Videnskaben og Kunstens Væsen” (“On the Nature of Science and Art”), in H. C. Ørsted, Samlede og Efterladte Skrifter, vol. 9 (Copenhagen, 1852), pp. 41–43.

KANT—NATURPHILOSOPHIE— ELECTROMAGNETISM MICHAEL FRIEDMAN

Robert Stauffer first gave prominence to the circ*mstance that Ørsted’s discovery of electromagnetism was intimately connected, at least in Ørsted’s own mind, with his rather deep philosophical engagement with both Kant’s philosophy of science, as presented in the Metaphysical Foundations of Natural Science of 1786, and the further development of this philosophy within post-Kantian German Naturphilosophie, especially as represented by Schelling.1 Building on Stauffer’s contribution, L. Pearce Williams developed a more general picture of the development of electromagnetic theory—especially in the work of Faraday but also placing particular emphasis on Ørsted—as resting on a profound deviation from the Newtonian tradition due to Naturphilosophie and ultimately to Kant.2 For Williams, it was Kant’s articulation of a so-called dynamical theory of matter in his treatise of 1786—a theory that represents matter as constituted out of the “fundamental forces” of attraction and repulsion rather than as a primitive, originally hard and impenetrable solid—which first opened up the possibility of a unified treatment of all the forces of nature, including, especially, the electric and magnetic forces.3 And Ørsted’s work, in particular, is then seen as a direct expression of this fundamentally Kantian view.4 Stauffer, for his part, makes a closely analogous argument to the effect that Ørsted’s “metaphysical faith” in “the unity of the powers of nature”—a faith nurtured and buttressed by his earlier philosophical work on Kant’s Metaphysical Foundations and German

1

2 3

4

Robert Stauffer, “Speculation and Experiment in the Background of Ørsted’s Discovery of Electromagnetism,” Isis 48 (1957), pp. 33–50. L. Pearce Williams, Michael Faraday: A Biography (New York: Basic Books, 1965). See Williams, Michael Faraday…, p. 62: “The reduction of all physical phenomena to attractive and repulsive forces was seductively simple. The different kinds of attraction and repulsion—electrical, magnetic, etc.—were the results of different conditions under which the two basic forces manifested themselves. Behind these differences lay the essential unity of all forces.” See Williams, Michael Faraday…, p. 137: “The possibility that electricity and magnetism were but different modes of action of the underlying and fundamental forces of attraction and repulsion was a primary tenet of the Kantian dynamic philosophy. One of Kant’s most ardent disciples, the young Dane Hans Christian Ørsted, devoted twenty years of his life to making this possibility manifest.”

135 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 135–158. © 2007 Springer.

136

M. FRIEDMAN

Naturphilosophie—played a key role in motivating and sustaining Ørsted’s later experimental work culminating in the discovery of electromagnetism.5 Yet these contributions of Stauffer and Pearce Williams, as important and original as they are, remain on an excessively high level of generality. For a commitment to a unified treatment of all the forces of nature—a faith in “the unity of the powers of nature”—is far too abstract and unspecific to carry very much explanatory weight as a factor in Ørsted’s quite particular discoveries. For example, Newton, in the Opticks, had already sketched a general program for developing a unified account of nature (including electrical and magnetic phenomena, of course) on the basis of a small number of attractive and repulsive “Powers” or “active Principles,” and it was precisely in virtue of this explanatory program, for Newton, that nature is “very comformable to herself and very simple.”6 Or, to take a more directly relevant example, when Helmholtz, in his famous monograph of 1847, provided the first theoretical basis for the principle of the conservation of the total quantity of force (i.e. energy) in all the processes of nature in which it is converted from one form into another (a principle that became emblematic, in the mid-19th-century, of just the kind of “unity of the powers of nature” that Stauffer and Pearce Williams have in mind),7 Helmholtz’s own theoretical framework employed paradigmatically Newtonian central forces acting both in straight lines and immediately at a distance. Indeed, Helmholtz, more generally, is quite critical of German Naturphilosophie and sees his own work, in this respect, as a return to the more serious and sober scientific preoccupations of the original Kant.8 Accordingly, while Helmholtz’s 1847 monograph is clearly indebted to Kant’s 1786 Metaphysical Foundations, it bears no traces at all of post-Kantian Naturphilosophie.

5

6

7

8

See Stauffer, “Speculation and Experiment…, p. 34. “To Ørsted, his belief in the existence of a physical relation between electricity and magnetism seemed a logical corollary to this general principle of unity [of all the powers of nature]. His metaphysical faith in this relationship afforded a motive for his persistent experimentation in the field of electricity and magnetism.” See Isaac Newton, Opticks (London: G. Bell and Sons, 1931), Query 31. “And thus Nature will be very comformable to her self and very simple, performing all the great Motions of the heavenly Bodies by the Attraction of Gravity which intercedes those Bodies, and almost all the small ones of their Particles by some other attractive and repelling Powers which intercede the Particles.” See, for example, the immediate continuation of the passage cited in note above: “From this it followed logically that all forces of nature were convertible into one another; one need only find the proper conditions for accomplishing the conversion.” The link between energy conservation and Naturphilosophie is also considered in a well-known paper by Thomas Kuhn, “Energy Conservation as an Example of Simultaneous Discovery,” in M. Clagett (ed.), Critical Problems in the History of Science (Madison: University of Wisconsin Press, 1959), which, with regard to electromagnetic forces, in particular, follows Stauffer’s account of Ørsted. It is for this reason that Helmholtz is standardly taken as an early representative of 19th-century neo-Kantianism—a movement aiming to turn away from the “metaphysical” speculations of postKantian idealism (as represented, in particular, by Schelling’s Naturphilosophie) to the original “epistemological” orientation of Kant himself. For discussion see the Introduction to Cassirer, The Problem of Knowledge: Philosophy, Science, and History since Hegel, translated by W. Woglom and C. Hendel (New Haven, CT: Yale University Press, 1950). Helmholtz’s celebrated address, “Über das Sehen des Menschen,” delivered at the dedication of a monument to Kant in Königsberg in 1855, condemns the then current rift between philosophy and natural science due, in Helmholtz’s opinion, to the entirely speculative Naturphilosophie of Schelling and Hegel, and announces a new project of cooperation between the two disciplines in the spirit of Kant. This address then became a model for the neo-Kantian tradition.

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

137

What is needed, then, is a more detailed account of Ørsted’s particular intellectual context, for it is precisely these details that distinguish Ørsted’s intellectual situation both from Newton’s (and also Kant’s) on the one side and from Helmholtz’s on the otherside. And one of the most important of these details, from the present point of view, is the circ*mstance that the dynamical theory of matter favored by both Ørsted in particular and Naturphilosophie more generally is fundamentally different from Kant’s original theory. Indeed, as I myself have just intimated, Kant’s theory, in most relevant respects, is far more Newtonian then theirs. It is especially misleading, therefore, when Pearce Williams portrays Kant as the great opponent of the Newtonian natural philosophy, for Kant’s philosophy of nature, in most relevant respects, should rather be viewed as a culmination of the Newtonian tradition.9 Whereas the dynamical theory of matter of the Metaphysical Foundations did serve, in fact, as a preamble or prolegomenon to the Naturphilosophie of Schelling and Ørsted, it is only by radically reconceptualizing and transforming Kant’s philosophy that the decisive step into Naturphilosophie was actually taken. Fortunately, the main outlines of a more detailed and specific account of Ørsted’s particular intellectual situation—both philosophically and scientifically—have already been provided by Andrew Wilson, in his outstanding Introduction to the Princeton edition of Ørsted’s (selected) scientific works.10 The principal points are these:

9

10

Thus, one central ambition of the dynamics of the Metaphysical Foundations is to defend the Newtonian attraction as a true and immediate action at a distance against both “all sophistries of a metaphysics that misunderstands itself ” (namely Leibnizean metaphysics) and Newton’s own doubts, which, according to Kant, “set [Newton] at variance with himself ” (Ak. 4, p. 514; the first quotation is at 523). Pearce Williams therefore commits a serious error when he portrays Kant’s transcendental idealism about space as fundamentally inimical to Newtonian action at a distance—see Pearce Williams, The Origins of Field Theory…, pp. 34–35): “Thus, The Critique of Pure Reason, if taken seriously as a fundamental critique of metaphysics, would mean the rejection out of hand of the very foundations of Newtonian physics. Particles could not act upon one another across empty space because empty space could not be known to the mind.” What Williams misses here is that, whereas Kant indeed denies that empty space can be an object of experience, he still insists, in the Metaphysical Foundations, that “[t]he attraction essential to all matter is an immediate action of matter on other matter through empty space” (Dynamics, Proposition 7, 512)—so that the action of gravitation, in particular, is entirely independent of any intervening matter that may fill the space between the attracting bodies (516): “this attraction is a penetrating force and acts immediately at a distance through that space, as an empty space, regardless of any matter lying in between.” Williams, “Kant, Naturphilosophie and Scientific Method,” in R. Giere and R. Westfall (eds.), Foundations of Scientific Method in the Nineteenth-Century (Bloomington, IN: Indiana University Press, 1973), pp. 18–19. acknowledges that he has deliberately “misread” Kant—“as the Naturphilosophen misread him.” However, while it is perfectly legitimate to depict Kant as the Naturphilosophen read him, it is even more important, I think, to become as clear as possible on the various ways in which they radically transformed him—and on the way in which, in particular, he is in fact much more Newtonian than they. See Jelved et al., Selected Scientific Writings of Hans Christian Ørsted…, pp. xv–xl. Another particularly helpful discussion is Gower, “Speculation in Physics.…,” although Gower himself paints a somewhat more skeptical picture of both Naturphilosophie in general and its relevance to Ørsted’s discovery in particular.

138

M. FRIEDMAN

1. Whereas Kant, in the Metaphysical Foundations, had declared that chemistry was not yet a science and therefore incapable of an a priori “metaphysical” foundation, one of the central ambitions of both Schelling and Ørsted was precisely to provide such a foundation. This was to proceed by further articulating the fundamental forces of attraction and repulsion Kant had shown to be constitutive of all matter in general and as such, so that chemical forces, in particular, would thereby naturally emerge. 2. New developments in electrochemistry provided the primary empirical impetus here, in so far as chemical forces could now be plausibly identified with electrical forces, and electrical forces, in turn, could be taken as a further articulation and development of the fundamental forces of attraction and repulsion in general. The electrochemical researches of Johann Ritter were particularly important in this regard, for Ritter had encountered Schelling’s philosophy at Jena and was also a close friend and collaborator of Ørsted’s; and it was Ritter, in fact, who first introduced Ørsted to this aspect of Schelling’s Naturphilosophie. 3. Galvanism or current electricity of course played a central role in the new electrochemistry, as exemplified, above all, in the electrolytic decomposition of water resulting in oxygen and hydrogen gasses. Oxygen and hydrogen were thereby associated with negative and positive electricity, respectively, and this suggested an especially close link between electrical forces and the fundamental chemical forces involved in combustion. Schelling, Ritter, and Ørsted (along with many others, of course) took this as evidence for the electrical nature of chemical affinities quite generally.11 4. Finally, the well-known parallels between electrical and magnetic forces suggested to Schelling, Ritter, and Ørsted that magnetism, too, was essentially implicated in chemical interactions (including galvanism) and, more specifically, that what Schelling called the basic or original form of “the dynamical process” was further differentiated, at the level immediately following that of Kant’s two fundamental forces of attraction and repulsion in general, into magnetism, electricity, and chemical forces (including galvanism). For Ørsted, this suggested a testable link between magnetism and galvanism, in particular, and thereby stimulated the experimental researches that eventually resulted in his great discovery. As Wilson admirably explains, the immediate intellectual context for Ørsted’s discovery therefore involved a very specific—and, as it turned out, especially timely—interaction between contemporary philosophical and scientific developments. Just when philosophy of nature in the Kantian tradition was ripe, as it were, for an extension of the dynamical theory of matter into chemistry, new discoveries in electrochemistry provided the needed fertile soil within which

11

This same work, of course, was the basis for Davy’s and Faraday’s parallel electrochemical researches in England. As Pearce Williams (among others) has made abundantly clear, these researches were also intimately involved with German Naturphilosophie, as mediated by the influence of Coleridge. But this specifically English context, together with Coleridge’s particular version of Naturphilosophie, lies well beyond the scope of our present discussion.

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

139

such a philosophy could then bear scientific fruit.12 What I here want to add to Wilson’s account is a bit more detail on the relationship between Kant’s dynamical theory of matter and Schelling’s, on the basis of which we might then be in a position to shed further light on both Ørsted’s philosophy of nature in particular and the relationship between philosophical speculation and scientific discovery more generally. Kant’s philosophy of human knowledge and experience—including his more specific philosophy of (corporeal) nature developed in the Metaphysical Foundations of Natural Science—is based on a number of fundamental distinctions. The most important is the distinction between the passive or receptive faculty of pure intuition or sensibility (involving the pure forms of sensible intuition, space and time) and the active or intellectual faculty of pure understanding (involving the categories or pure concepts of the understanding: substance, causality, community, and so on). It is this distinction that gives rise to the dichotomy between appearances (spatio-temporal objects given to our sensibility) and things in themselves (purely intellectual objects thought by the understanding alone), as well as the closely related distinction between constitutive a priori principles and merely regulative a priori principles. Constitutive principles result from the application of purely intellectual representations to our spatio-temporal sensibility and yield necessary conditions for all objects of experience—conditions which therefore are necessarily realized in experience. The pure concepts or categories of substance, causality, and community, for example, are necessarily realized in our experience by a system of causally interacting conserved entities distributed in space and time—a system for which massive bodies interacting in accordance with Newtonian universal gravitation acting immediately at a distance consistently provided Kant with his primary model. Indeed, one of the main points of the Metaphysical Foundations is to explain how the general constitutive principles of experience of the Critique of Pure Reason are further specified or articulated to provide an a priori “metaphysical” foundation for precisely this Newtonian model.13 Yet Kant’s general constitutive grounding of experience, even as extended into “the special metaphysical of corporeal nature” in the Metaphysical Foundations, leaves much of the natural world still unaccounted for. The Metaphysical Foundations only provides a priori insight into the most general properties and powers of all matter in general and as such (properties such as mass, gravity, impenetrability, and elasticity) and leaves even the property of cohesion for a physical and

12

13

Wilson also uses this quite specific intellectual situation, in the context of a detailed discussion of the development of Ørsted’s electrochemical experimental program, to provide a fundamental clarification of Ørsted’s theoretical ideas on the nature of electricity and magnetism—including the famous conflictus—a topic that I will have to leave aside here. This idea is not particularly controversial today. For discussion Friedman, Kant and the Exact Sciences (Cambridge, MA: Harvard University Press, 1992)—although particular details of my interpretation are, of course, controversial. Compare note above.

140

M. FRIEDMAN

empirical rather than an a priori and metaphysical treatment.14 Accordingly, Kant here assigns the problem of further specifying the general concept of matter as such into particular species and subspecies—the problem of what he calls “the specific variety of matter”—for the further development of empirical physics and chemistry; and so it is by no means surprising that Kant, as we have said, officially denies scientific status to chemistry and asserts, in the Preface to the Metaphysical Foundations, that, at least at present, “chemistry can be nothing more than a systematic art or experimental doctrine, but never a proper science.”15 How, then, are the more empirical sciences of nature to proceed? It is here that Kant invokes his famous doctrine of the regulative use of reason. The faculty of reason, in contradistinction to the faculty of understanding, generates a priori intellectual representations that cannot be fully realized in our human spatio-temporal experience. These include the ideas of God, Freedom, and Immortality, for example, and also, more relevant to our present concerns, the idea of the systematic unity of all empirical concepts and principles under the a priori constitutive concepts and principles already generated by the understanding. This idea of systematic unity guides our process of inquiry in the more empirical and inductive sciences, without constitutively constraining it, as we successively ascend from lower level empirical concepts and principles toward higher level concepts and principles. The goal of this process is an ideal complete empirical science of nature in which all empirical concepts and principles are constitutively grounded in the pure categories and principles of the understanding, but this is necessarily an ideal we can only successively approximate but never actually attain. And the paradigmatic application of the regulative use of reason, in the period of both the Metaphysical Foundations and the first Critique, is precisely to contemporary chemistry. Kant sees this chemistry—primarily Stahlian phlogistic chemistry as supplemented by the new discoveries in pneumatics but not yet including Lavoisier—as a purely empirical or experimental art guided by the regulative use of reason towards an entirely unspecified and indeterminate future state of affairs in which the experimental results in question are finally grounded in the fundamental forces of matter in a way that we are not yet in a position to anticipate.16 Moreover, as is well known, Kant, in the Critique of Judgement, extends the doctrine of the regulative use of reason to what he now calls the faculty of

14

15

16

See (Ak. 4, p. 518): “The action of the universal attraction immediately exerted by each matter on all matters, and at all distances, is called gravitation; the tendency to move in the direction of greater gravitation is weight. The action of the general repulsive force of the parts of every given matter is called its original elasticity. Hence this property and weight constitute the sole universal characteristics of matter, which are comprehensible a priori, the former internally, and the latter in external relations. For the possibility of matter itself rests on these two properties. Cohesion, if this is explicated as the mutual attraction of matter limited solely to the condition of contact, does not belong to the possibility of matter in general, and cannot therefore be cognized a priori as bound up with this. This property would therefore not be metaphysical but rather physical, and so would not belong to our present considerations.” Immanuel Kant, Metaphysical Foundations of Natural Science, translated by M. Friedman. In H. Allison and P. Heath (eds.), Immanuel Kant: Theoretical Philosophy after 1781 (Cambridge: Cambridge University Press, 2002), Ak. 4, p. 471. For further discussion of Kant’s conception of chemistry in this period see Friedman (1992, ch. 5).

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

141

reflective judgement, and he now applies this faculty, in particular, to the case of biology. The problem here, in a nutshell, is that all matter in general and as such—all matter as the object of our outer senses in space—is essentially lifeless. This, in fact, is how Kant interprets the law of inertia, which law, in turn, is itself constitutively grounded by a further specification of the a priori principle of causality articulated in the first Critique.17 Biology, the study of life, can therefore never be a science in the strict sense for Kant; it can never be constitutively grounded in the fundamental forces of matter. The best we can do, in this case, is to extend the doctrine of the regulative use of reason via the teleological idea of purposiveness [Zweckmäßigkeit]—an idea which already arises for reflective judgement in general as it guides our inductive ascent from particular to universal toward the ideal infinitely distant goal of a complete systematic unity of nature. And this idea can now be applied to particular objects of nature or “natural products” (i.e. living organisms) in so far as they are conceived, by reflective judgement, as themselves purposively organized. But such a mode of conception is in no way constitutive of these objects themselves; it is rather a merely regulative device for guiding our empirical inquiry into living organisms as far as it may proceed. Post-Kantian German idealism—as successively articulated by Fichte, Schelling, and Hegel—is characterized by a rejection of Kant’s fundamental distinctions, a rejection, that is, of “Kantian dualism.” A prominent example of such dualism, of course, is the distinction between appearances and things in themselves and, more relevant to our present concerns, the closely related distinction between constitutive and regulative principles. In particular, it appears that Kant’s doctrine of the regulative use of reason inevitably leaves us with a quite intolerable skepticism concerning most of the phenomena of nature. For only very few of these phenomena, as we have seen, are actually constitutively grounded, and, for the rest, we have at best the otherwise entirely indeterminate hope that they might be constitutively grounded some day. (In the case of biology, as we have seen, even this much hope seems to be in vain.) It would appear, then, that the vast majority of natural phenomena are not objectively (constitutively) grounded at all, and our claims to have rational or objective knowledge of nature are accordingly cast into doubt.18 For the post-Kantian idealists, therefore, the

17

18

The statement of the law of inertia and accompanying proof is Proposition 3 of the Mechanics chapter of the Metaphysical Foundations. The following Remark explains the connection with lifelessness—namely, the nonexistence of any internal principle of change. See, in particular, the conclusion of this Remark (Ak. 4, p. 544): “The possibility of a proper natural science rests entirely and completely on the law of inertia (along with that of the persistence of substance). The opposite of this, and thus also the death of all natural philosophy, would be hylozoism. From this very same concept of inertia, as mere lifelessness, it follows at once that it does not mean a positive striving to conserve its state. Only living beings are called inert in this latter sense, because they have a representation of another state, which they abhor, and against which they exert their power.” This rejection of “hylozoism” is clearly directed at Leibnizean natural philosophy. My formulation of this skeptical problem is indebted to Paul Franks (ed.), “What should Kantians learn from Maimon’s Skepticism?,” in G. Freudenthal, ed, The Philosophy of Salomon Maimon and its Place in the Enlightenment (Dordrecht The Netherlands: Kluwer, forthcoming), although Franks himself does not emphasize, as I do, the distinction between constitutive and regulative principles—he instead formulates what I take to be essentially the same problem by means of a distinction between scientific judgments and everyday or “ordinary” judgments.

142

M. FRIEDMAN

very enterprise of transcendental philosophy—the attempt to give an a priori or rational foundation for the totality of our knowledge and experience—must be radically reconceived. This radical reconceptualization of transcendental philosophy proceeds in outline as follows. Since, to begin with, we are also rejecting Kant’s central distinction between a passive or receptive faculty of pure sensibility and an active or intellectual faculty of pure understanding, we must reject Kant’s version of the distinction between understanding and reason as well; for the understanding, according to Kant, is the intellectual faculty applied to sensibility, whereas reason, in contradistinction to the understanding, is the same intellectual faculty considered independently of sensibility.19 Here it issues in the ideas of God, Freedom, and Immortality and also, as we have seen, in the regulative use of reason under the idea of systematic unity: an infinite progressive process which approximates, but never fully attains, an infinitely distant goal. Moreover, this process, for Kant, also has an essentially “dialectical” character, in that it inevitably gives rise to the antinomies of pure reason (the idea of the world as either finite or infinite, and so on), which can only be resolved by a higher synthesis based on a thoroughgoing appreciation of the essentially incompletable nature of the dialectical process as such. The faculty of reason, considered in itself, is thus both infinitary and dialectical; and it is therefore very natural, once we have abandoned, once and for all, the distinction between active intellect and passive sensibility, to view the a priori constitution of knowledge and experience as an infinitary dialectical progression of the same kind.20 But how is such an infinitary dialectical progress of reason, which, for Kant, is a reflection of how we must (regulatively) view the objects of nature rather than how they (constitutively) are, to be reconceived as an a priori constitution of nature itself ? It is precisely here that Schelling makes his decisive contribution. For Schelling, 19

20

It does not follow, as Allen Wood has emphasized to me, that the post-Kantian idealists have no room for reinterpreting the distinction between understanding and reason in other terms—indeed, new versions of this distinction become central to their thought. The main point here, however, is that they have absolutely no room for Kant’s particular version, which entails, in particular, a sharp distinction between constitutive principles (due to the understanding) and merely regulative principles (due to reason). For Kant, by contrast, the constitutive grounding of experience by the understanding is necessarily finite (since there are only twelve categories), and only the contrasting regulative use of reason is infinite. Kant makes precisely this point, in fact, in the Preface to the Metaphysical Foundations (Ak. 4, p. 473): “[J]ust as in the metaphysics of nature in general, here also the completeness of the metaphysics of corporeal nature can confidently be expected. The reason is that in metaphysics the object is only considered in accordance with the general laws of thought, whereas in other sciences it must be represented in accordance with data of intuition (pure as well as empirical), where the former, because here the object has to be compared always with all the necessary laws of thought, must yield a determinate number of cognitions that may be completely exhausted, but the latter, because they offer an infinite manifold of intuitions (pure or empirical), and thus an infinite manifold of objects of thought, never attain absolute completeness, but can always be extended to infinity, as in pure mathematics and empirical doctrine of nature.” For helpful discussions of postKantian idealism along these lines see, e.g. Royce, Lectures on Modern Idealism (New Haven, CT: Yale University Press, 1919), and, for a more advanced treatment, Cassirer, Das Erkenntnisproblem in der Philosophie und Wissenschaft der neureren Zeit. Dritter Band: Die Nachkantischen Systeme (Berlin: Bruno Cassirer, 1920). For an important recent discussion, placing particular emphasis on the role of Naturphilosophie and the resulting “organic view of nature,” see Beiser, German Idealism (Cambridge, MA: Harvard University Press, 2002).

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

143

transcendental philosophy, the story of how human reason successively approximates to a more and more adequate picture of nature, has a necessary counterpart or dual, as it were, in Naturphilosophie, the story of how nature itself successively unfolds or dialectically evolves from the “dead” or inert matter considered in statics and mechanics, to the essentially dynamical forms of interaction considered in chemistry, and finally to the living or organic matter considered in biology. Since nature, on this view, dialectically unfolds or successively evolves in a way that precisely mirrors the evolution or development of our rational conception of nature (and, of course, vice versa), it follows that there is no possible skeptical gap between nature itself and our conception of it, or, in Kantian terminology, between the constitutive domain of the understanding and the merely regulative domain of reason. All the phenomena of nature—including, in particular, both chemical and biological phenomena—are rationally or objectively grounded in the same way. The key to Schelling’s conception is a dialectical extension and elaboration of Kant’s original dynamical theory of matter. From Schelling’s point of view, Kant’s dynamical theory of the fundamental forces of attraction and repulsion necessary to all matter in general and as such (which embraces, therefore, even the “dead” or inert matter considered in statics and mechanics) has already introduced an essentially dialectical element into nature, insofar as the dynamical constitution of matter in general proceeds from the positive reality of expansive force (repulsion), through the negative reality of contractive force (attraction), to the limitation or balance of the two in a state of equilibrium.21 But we now know, as Kant himself did not, that chemistry can be dynamically grounded by a dialectical continuation of this progression—as we proceed, more specifically, from the magnetic, through the electrical, to the chemical (or galvanic) forms of the basic or original dynamical process grounded in the fundamental forces of attraction and repulsion. And, once we have gone this far, it is then a very short step (particularly in view of the recently discovered parallel interconnections among electrical, galvanic, and biological phenomena) to view biology, too, as a further dialectical continuation of the same dynamical process.22 Biology, too, can be a science, for all rational science, as Kant did not see, is grounded in a single dynamical evolutionary dialectical progression. The whole of nature, in this sense, is at once both rational and

21

22

This already represents a radical transformation and reinterpretation of Kant’s original dynamical theory. To be sure, Kant’s construction of matter in general out of the two fundamental forces of attraction and repulsion proceeds via the three categories of quality: reality, negation, and limitation. For Kant himself, however, in the necessarily finite constitutive grounding of experience provided by his metaphysics of nature (see note above), these categories of quality, in turn, are necessarily subordinate to the categories of relation articulated in what he calls the analogies of experience: the categories of substance, causality, and community. In the Metaphysical Foundations these latter are realized by the laws of conservation of matter, inertia, and equality of action and reaction—which laws, as Kant understands them, entail the lifelessness of matter in general (see notes). This overriding emphasis on what he calls the three “laws of mechanics” is central to Kant’s much more Newtonian version of the dynamical theory (see notes above). For Schelling, this dialectical continuation takes a quite precise and specific form: corresponding to the magnetic, electrical, and chemical (or galvanic) forms, we then have, as the “third potency” of the original dynamical process, the biological powers of reproduction, (nervous) irritability, and sensibility (see Ideas for a Philosophy of Nature, Supplement to Book I, ch. 6).

144

M. FRIEDMAN

alive23; and this means, in particular, that there actually is life—objectively, not merely regulatively—in even the very simplest forms of organized matter.24 At the center of this entrancing vision stands the new electrochemistry. For it is precisely here that we can unite the concept of matter in general as conceived in Kant’s original dynamical theory (the “dead” matter considered in statics and mechanics) with matter as conceived by Naturphilosophie—as an inexhaustible source of rational life. It is in precisely this context that we can view chemistry, in Schelling’s words, as a dialectical “middle term” between mechanism, on the one side, and biological (ultimately rational) living purposiveness, on the other.25 Indeed, even the inert matter considered in statics and mechanics is already at least potentially alive, since Kant’s dynamical theory has shown that the fundamental forces of attraction and repulsion are necessary to all matter in general and as such, and we have just seen that the original or primary dynamical process governed by these forces must necessarily evolve or develop into first chemical and then biological forms of external nature. In the end, it is precisely by rejecting the fundamental Kantian contention that all matter in general and as such—all matter as the object of our outer senses—is essentially lifeless that Schelling, from his point of view, finally overcomes any possibility of a skeptical gap between our rational conception of nature and nature itself.26

23

24

25

26

See Ideas for a Philosophy of Nature, translated by E. Harris and P. Heath (Cambridge: Cambridge University Press, 1988), p. 40: “Finally, if we comprehend nature as a single whole, then mechanism, i.e. a past-directed series of causes and effects, and purposiveness [Zweckmäßigkeit], i.e. independence of mechanism, simultaneity of causes and effects, stand opposed to one another. In so far as we now unite these two extremes, an idea of a purposiveness of the whole arises in us—nature becomes a circle that returns into itself, a self-enclosed system.” See Ideas…, p. 35: “This philosophy must admit, therefore, that there is a graduated development [Stufenfolge] of life in nature. Even in mere organized matter there is life, but only life of a limited kind. This idea is so old, and has been preserved until now in the most varied forms up to the present day—(already in the most ancient times the whole world was [regarded as] penetrated by a living principle, called the world-soul, and Leibniz’s later period gave every plant its soul)—that one can well surmise in advance that some ground for this natural belief must lie in the human spirit itself. And it is in fact so. The entire mystery surrounding the problem of the origin of organized bodies rests on the circ*mstance that in these things necessity and contingency are united in the most intimate way. Necessity, because their existence is already purposive, not only (as in the case of the work of art) their form; contingency, because this purposiveness is nonetheless only actual for an intuiting and reflecting being.” See Ideas…, p. 149: “Therefore, already in the chemical properties of matter there actually lie the first, although still completely undeveloped seeds of a future system of nature, which can unfold into the most varied forms and structures, up to the point where creative nature appears to return back into itself. Thus, at the same time, further investigations are marked out, up to the point where the necessary and the contingent, the mechanical and the free, separate from one another. Chemical phenomena constitute the middle term between the two. It is this far, then, that the principles of attraction and repulsion actually lead, as soon as one considers them as principles of a universal system of nature.” Schelling is here self-consciously returning to precisely the Leibnizean “hylozoism” Kant explicitly rejects: compare notes, and above. This essentially biological or organic conception of nature then implies the overcoming of all skepticism in the sense that the closing of the circle mentioned above (embracing both mechanism and teleology) means that transcendental philosophy and Naturphilosophie—spirit and nature—are ultimately identical, insofar as nature itself gives rise to both life in general and conscious or rational life in particular. See Ideas…, p. 42: “Nature should be the visible spirit, spirit the invisible nature. It is here, therefore, in the absolute identity of spirit within us and nature outside us, that the problem of how a nature outside us is possible must be solved.”

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

145

It is also clear, as Wilson explains (note above), that this grand naturphilosophisch vision constituted an essential part of Ørsted’s intellectual development. Ørsted began his career as a philosopher of nature with two dissertations on Kant’s Metaphysical Foundations in 1799; yet, even here, his aim, unlike Kant’s, was to secure an a priori “metaphysical” foundation for matter-theory and chemistry, including the fundamental property of cohesion.27 At this time, however, Ørsted was not yet acquainted with either the new developments in electrochemistry or with Schelling’s radical transformation of Kant’s original dynamical theory of matter intended to account for them.28 When he did become acquainted with these developments soon thereafter, Ørsted enthusiastically embraced Schelling’s key idea of chemistry as an extension of general dynamics29; and he came to share Schelling’s view, in particular, of how magnetic, electrical, and chemical (or galvanic) phenomena constitute what Schelling calls the “second potency” of the original dynamical process.30 He also came to share, more generally, the broader philosophical vision of a dialectical or evolutionary development of nature as a whole in the realization or progressive unfolding of a single infinite rational life:

27

28

29

30

Compare note above. Accordingly, Ørsted here envisions a far-reaching alteration of the very structure of the Metaphysical Foundations, so as to include a theory of motion on the one side and a (separate) theory of matter on the other. This already represents a fundamental (and quite self-conscious) divergence from Kant’s own procedure, according to which “natural science [throughout] is either a pure or applied doctrine of motion” (Ak. pp. 4, 476). Ørsted does refer to he first edition of Schelling’s Ideas, published in 1797. But this, of course, was prior to the invention of the Voltaic pile and the consequent discovery of the electrolytic decomposition of water by Nicholson and Carlisle in 1800. Schelling himself had already become convinced of a close link between electrical forces and chemical forces (especially combustion) on the basis of electrochemical phenomena involving static electricity (sparking)—including the researches of Priestley and Cavendish, as well as early work on the decomposition of water by sparking of the Dutch chemists Van Troostwijk and Deiman (see Ideas, Book I, ch. 4). It appears, as already suggested, that Ørsted was initiated into much of this new material by Ritter: see, e.g. Christensen, “The Ørsted-Ritter Partnership and the Birth of Romantic Natural Philosophy,” Annals of Science 52 (1995), pp.153–185. Ritter himself performed very early experiments with the Voltaic pile and, in particular, discovered the electrolytic decomposition of water independently of Nicholson and Carlisle. (Ritter also originated the electrochemical series linking the affinities of metals for oxygen with the metals employed in the Voltaic pile.). Schelling reports on this work of Ritter’s in the second, 1803 edition of the Ideas (Supplement to Book I, ch. 3), and it was this edition, in particular, which then exerted a decisive influence on the further development of Ørsted’s philosophy of nature. Again, Schelling is much clearer about this crucial piece of the story in the second, 1803 edition. On this view, magnetic forces present the basic or elementary form of the cohesion of bodies, expressed in a one-dimensional line (between two magnetic poles), electrical forces act at the two-dimensional surfaces of bodies determined by such cohesion (in a distribution of charge manifesting static electricity), and chemical or galvanic forces act within the three-dimensional volumes of such bodies (as their parts move relative to one another under the influence of affinities and other chemical processes). See Ideas …, pp. 115–118, 128–129, 137, 180, 204–205, 219, 268–271. Compare Ørsted, Naturvidenskabelige Skrifter (Scientific Papers), 3 vols. (Copenhagen: Royal Danish Society of Sciences, 1920), vol. 3, pp. 103–105; vol. 3, 110; vol. 3, 113–116; vol. 1, 330–332; vol. 3, 165–166; vol. 2, 39–40; vol. 2, 146–149), Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, pp. 190–191, 195, 197–199, 252–253, 291, 312, 378–789). For a detailed account of how Ørsted’s theoretical and experimental work takes off from the second edition of Schelling’s Ideas, culminating in the grand naturphilosophisch vision considered immediately below, see Wilson’s discussion in Jelved et al. (1998, xxix–xxxix).

146

M. FRIEDMAN A clearer perspective soon adds to this that there is nothing dead and rigid in nature, but that every thing exists only as a result of a development [Entwicklung], that this development proceeds according to laws, and that, therefore, the essence of every thing is based on the totality of the laws or on the unity of the laws, i.e., the higher law by which it has been brought forth. Every thing, however, must again be regarded as an active organ of a more comprehensive whole, which again belongs to a higher whole so that only the great All sets the limit of this progression. And thus the universe itself would be regarded as the totality of the developments, and its law would be the unity of all other laws. However, what finally gives the study of nature its highest meaning is the clear understanding that natural laws are identical to the laws of reason, so they are in their application identical to thoughts; the totality of the laws of a thing, regarded as its essence, is therefore an idea of nature, and the law or the essence of the universe is the totality [Inbegriff ] of all ideas, identical with absolute reason. And so we see all of nature as the appearance [Erscheinung] of one infinite force and one infinite reason united, as the revelation of God.31

As the surrounding context makes clear, a primary motivation for this broader vision is precisely the new discoveries in electrochemistry, as interpreted, along with Schelling, as expressing the fundamental unity of magnetic, electrical, and galvanic forces.32 It is with perfect justice, therefore, that Ørsted himself describes his discovery of electromagnetism, in the now well-known passage with which Stauffer begins his 1957 paper, by asserting that he “was not so much led to this [i.e. the view that ‘magnetical effects are produced by the same powers as the electrical’] by the reasons commonly alleged for this opinion, as by the philosophical principle, that all phenomena are produced by the same original power.”33 We can deepen our appreciation of the quite specific intellectual situation faced by Schelling, Ritter, and Ørsted by taking a brief look at how Kant himself, very late in his career, attempted to extend his dynamical theory of matter into chemistry. This attempt occurs in unpublished materials from the years 1796–1803 collected into what is now known as the Opus postumum, and it involves, in Kant’s own terminology, drafts and sketches of a projected new work to be entitled Transition from the Metaphysical Foundations of Natural Science to Physics. Here we can see Kant responding to two complementary pressures. On the one hand, there is a clear systematic need to have something more to say about how the general concept of matter as such is further specified, a priori, into various species and subspecies—the problem, that is, of what Kant calls “the specific variety of matter.”

31

32

33

Jelved et al., Selected Scientific Works of Hans Christian, p. 384; translation emended), Ørsted, Naturvidenskabelige Skrifter…, vol. 2, 156–157). This occurs near the conclusion of the 1812 paper on “The Chemical Laws of Nature Obtained Through Recent Discoveries.” Compare again note above. It is several pages earlier in his 1812 paper that Ørsted famously suggests that “[t]he mode of action [Wirkungsform] in the circuit, or the galvanic [mode], stands between the pure electrical and the magnetic [modes] in that its forces are bound far more than in the former and far less than in the latter,” so that “one should experimentally investigate [versuchen] whether one can produce some effect on the magnet as a magnet in one of the states in which electricity occurs as very bound [i.e. by galvanic or current electricity].” Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, pp. 378–379; translation emended); Ørsted, Naturvidenskabelige Skrifter.…, vol. 2, pp. 147–148). Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, p. 546; Ørsted, Naturvidenskabelige Skrifter…, vol. 2, p. 356. Immediately thereafter Ørsted refers to the passage from 1812 cited in note above. Skeptical doubts about Ørsted’s later portrayal of his discovery—as expressed, for example, in Gower, “Speculation in Physics”…, pp. 345–346)—are therefore (in view of what we have just seen) quite unwarranted.

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

147

In the Metaphysical Foundations, as we have seen, Kant had very little to say about this, although he does present an outline of his conception of the foundations of contemporary chemistry (including such topics as cohesion, fluidity, elasticity, and dissolution).34 On the other hand, however, Kant, by 1795, has completed the transition from the phlogistic chemistry of Stahl to the new chemistry of Lavoisier, and he has now become convinced, contrary to the doctrine of the Metaphysical Foundations, that chemistry has finally entered the secure path of a science after all. There is a clear need, therefore, to explain the a priori foundations of this new science; and the Transition project, accordingly, is best conceived as an attempt further to develop transcendental philosophy so as to solve both of these problems at the same time—to provide an a priori explanation of the possibility of the specific variety of matter that is simultaneously an a priori grounding of Lavoisier’s new chemistry.35 The Transition project begins, naturally enough, from the idea that the two fundamental forces of attraction and repulsion constitutive of all matter in general and as such must be somehow further specified so as to account, at least in outline, for the more specific forces and powers exerted by specific types of matter—matter in the solid, liquid, or gaseous states, for example, or matter in the form of particular kinds of chemical substances. The Preface to the Metaphysical Foundations had asserted that chemistry could not become a science properly speaking until forces of attraction and repulsion appropriate to specifically chemical interactions could actually be discovered (forces, for example, of the kind Newton speculates about in the Opticks); and, since such forces had not yet been found, chemistry was not yet a science.36 Now, however, chemistry has become a science, but it has not done so, in the work of Lavoisier, by finally uncovering the specifically chemical forces. What Lavoisier has achieved, rather, is a fundamental conceptual reorganization of the subject—a new system of chemical classification based on the newly discovered central role of oxygen. The task of the Transition

34

35

36

This discussion occurs in the General Remark to Dynamics. Kant here introduces it as follows (Ak. 4, p. 525): “Instead of a sufficient explanation of the possibility of matter and its specific variety on the basis of these fundamental forces [of attraction and repulsion], which I cannot provide, I will present completely, so I hope, the moments to which its specific variety must collectively be reducible (although even so not conceivable in regard to its possibility).” For further discussion of the evolution of Kant’s conception of chemistry between the years 1786 and 1795, together with a reading of the Transition project along these general lines, see again Friedman, Kant and the Exact Sciences…, ch. 5. See Metaphysical Foundations (Ak. 4, pp. 470–471): “So long, therefore, as there is still for chemical actions of matters on one another no concept to be discovered that can be constructed, that is, no law of the approach or withdrawal of the parts of matter can be specified according to which, perhaps in proportion to their density or the like, their motions and all the consequences thereof can be made intuitive and presented a priori in space (a demand that will only with great difficulty ever be fulfilled), then chemistry can be nothing more than a systematic art or experimental doctrine, but never a proper science, because its principles are merely empirical, and allow of no a priori presentation in intuition. Consequently, they do not in the least make the principles of chemical appearances conceivable with respect to their possibility, for they are not receptive to the application of mathematics.”

148

M. FRIEDMAN

project, accordingly, is to show how this system of chemical classification itself receives an a priori grounding in transcendental philosophy. This task, as one might imagine, is not an easy one, and Kant struggles mightily with it throughout a large and bewildering series of outlines, drafts, and sketches of his projected new work. Finally, in 1799, in drafts that came closer than any other part of the Opus postumum to actual publication, he presents the so-called Aether Deduction: an a priori proof that there is a universally distributed aether or caloric fluid, constituted by a perpetual interaction between the two fundamental forces of attraction and repulsion, filling all of space. This universally distributed aetherial medium is supposed to provide an a priori grounding for the central theoretical construct of Lavoisier’s new chemistry—caloric or the imponderable matter of heat—and, at the same time, to serve, in a way that had long been familiar in 18th-century matter theory, as the medium or vehicle for light, electricity, and magnetism as well.37 In this way, the totality of forces or powers of nature—including, above all, the specifically chemical forces—are, at least in principle, systematically unified in a single a priori representation, and, in this sense, the problem of the specific variety of matter has finally been solved. The a priori representation in question—the representation of a universally distributed caloric fluid or aetherial medium filling all of space—is, by the standards of Kant’s critical philosophy, an extremely peculiar one. In particular, it combines aspects of discursive or conceptual representation (what Kant calls “distributive” or “analytic universality”) with the apparently entirely opposed characteristics of sensible or intuitive representation (what Kant calls “collective” or “synthetic universality”). As a continuum of forces providing a basis for further specification of the concept of matter in general, it is a discursive or conceptual representation; as a space-filling continuum, providing what Kant calls a perceptual “realization” of the pure intuition of space, it is a sensible or intuitive representation. Moreover, and by the same token, as an a priori principle for the further specification of the concept of matter in general it is a constitutive representation; as the ultimate ground for the systematic unity of all of the forces of matter it is a regulative representation. It is in this way, in fact, that the “top-down” constitutive procedure of the Metaphysical Foundations and the first Critique has a necessary intersection, as it were, with the “bottom-up” regulative procedure of the faculty of reflective judgement; and it is in precisely this way, accordingly, that the skeptical problems arising from the doctrine of the regulative use of reason that so vexed the postKantian idealists are finally resolved for Kant himself.38 37

38

Kant had consistently believed in the existence of such an aetherial medium throughout his long career (including, in particular, in the period of the first Critique and the Metaphysical Foundations). What is new, in the Opus postumum, is the idea that the existence of this medium can now be proved a priori. The central role of the matter of heat in Lavoisier’s chemistry provided Kant with new grounds for optimism that such a proof is (and must be) possible. Compare the paragraph to which note above is appended together with the preceding paragraph. This reading of the aether-deduction, as developed in Friedman, Kant and the Exact Sciences…, ch. 5, § IV), is quite controversial. For an alternative discussion, including some criticisms of my reading, see Förster, Kant’s Final Synthesis (Cambridge: Cambridge University Press, 2000). Another important aspect of the Opus postumum, to which Förster pays particular attention, is a reconceptualization of the place of biology and teleology in the critical system.

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

149

We cannot delve more deeply into the Opus postumum here. But we have seen enough, I hope, to appreciate that Kant himself, at the very end of his career, was right on the verge (both philosophically and scientifically) of the radically new conceptual situation faced by Schelling, Ritter, and Ørsted. He had already considered the problem of extending the dynamical theory of matter into chemistry, and, at the same time, he had already subjected the fundamental distinction between constitutive principles and regulative principles to a radical reconceptualization. Nevertheless, the decisive step into Naturphilosophie is one that Kant did not and could not take. For, in the first place, the critical new developments in magnetism, electricity, and chemistry that provided the fertile empirical soil on which alone Naturphilosophie could take root played no role at all in Kant’s thought. Kant, to the best of my knowledge, never engaged with even the electrostatic and magnetostatic work of Coulomb, to say nothing of the electrochemical researches arising from the Voltaic pile. The central idea of Naturphilosophie in its application to chemistry—that chemical forces are at bottom electrical in nature—never occurred to him; and, as a result, the idea that one could extend the dynamical theory of matter by conceiving magnetic, electrical, and galvanic forces as a further dialectical development of the original dynamical process governing the fundamental forces of attraction and repulsion was entirely foreign to Kant’s own final attempt to solve the problem of the specific variety of matter.39 Moreover, and in the second place, although, as we have seen, there may be room, with considerable stretching and straining, to find a place in Kant’s critical system for a representation that combines both constitutive and regulative aspects, there is no room at all for the grand naturphilosophisch vision of nature as a whole as the realization or evolutionary development of a single divine infinite rational life. For this idea, of course, entails the total Aufhebung of all of Kant’s most fundamental distinctions, along with the critical philosophy itself.40 It is in this precise sense that Kant’s own version of the dynamical theory of matter represents the prolegomenon—but only the prolegomenon—to the radically transformed version of the theory adopted by Schelling, Ritter, and Ørsted. Seeing how close Kant actually came to their particular intellectual situation—but also how far he remained—helps us develop a better understanding of the extraordinary, and extraordinarily timely, confluence of scientific problems and philosophical ideas which framed their enterprise and thereby made it fruitful. Our discussion of the relationship between Kant’s dynamical theory of matter and Schelling’s has tended to support, with a bit more specificity and detail, the original evaluations of Ørsted’s philosophical inspiration contributed by Stauffer and Pearce Williams, according to which Schelling’s Naturphilosophie, in particular,

39

40

By contrast, precisely this idea, as we have seen, lies at the heart of Schelling’s approach. See Ideas…, p. 205): “[A]ll particular or specific determinations of matter have their basis in the differing relation of bodies to magnetism, electricity, and the chemical process.” The representation of God—of an infinite rational being—is paradigmatic, for Kant, of a thing in itself lying entirely beyond all human knowledge, at least from a purely theoretical point of view. For the post-Kantian idealists, by contrast, the very distinction between theoretical and practical points of view is also necessarily called into question.

150

M. FRIEDMAN

provided essential philosophical motivation for Ørsted’s scientific discoveries. But these evaluations, not surprisingly, have since been challenged—first in a relatively measured and balanced way by Barry Gower and then in a very pointed and more contentious form by Timothy Shanahan.41 The basis for Gower’s skepticism is the conviction that Schelling’s metaphysics is simply too abstruse and obscure to have exerted a substantial influence on the actual practice of science42; whereas the focus of Shanahan’s attack, by contrast, is the idea that Schelling’s speculative methodology is too rationalistic and aprioristic vis-à-vis empirical physics—for Shanahan, accordingly, the more sober and empirical approach of Kant’s original dynamical theory, quite independently of its further elaboration by Schelling, provided, by itself, the decisive philosophical inspiration.43 A consideration of these challenges, by way of conclusion, will, I think, prove useful, not so much for the sake of deciding whether Schelling’s philosophy was or was not scientifically important, but rather as a means for shedding further light on the question of how philosophical speculation fruitfully interacts with scientific discovery more generally. Let us begin with the more focused and pointed challenge raised by Shanahan, which raises empirically oriented doubts concerning Schelling’s aprioristic methodology. In this connection, especially, it is appropriate to consider Ørsted’s own methodological remarks; and here it is clear, throughout his intellectual career, that Ørsted’s own view was that there is a necessary correspondence or parallelism between a purely aprioristic speculative deduction of natural phenomena on the one side and an empirical or experimental inquiry into these same phenomena on the otherside. This view is expressed, in the first instance, in Ørsted’s

41

42

43

Barry Gower, “Speculation in Physics: The History and Practice of Naturphilosophie,” Studies in the History and Philosophy of Science 3 (1973), pp. 301–356; Timothy Shanahan, “Kant, Naturphilosophie, and Ørsted’s Discovery of Electromagnetism: A Reassessment,” Studies in the History and Philosophy of Science 20 (1989), pp. 387–305. See, e.g. Gower, “Speculation in Physics…, p. 302: “This preoccupation with metaphysics provides a reason for skepticism concerning the impact of Naturphilosophie upon early nineteenth-century science. It is difficult to believe that anyone whose energies were primarily devoted to science could have employed with any confidence the immensely complex metaphysics of Schelling and others. The idiom in which the philosophical spokesmen for the movement chose to express their views is notorious for its virtually impenetrable abstractness. To this extent, Naturphilosophie is not comparable with, say, Cartesianism or Leibnizeanism where the metaphysical component is recognizably relevant to the development of science.” See, e.g. Shanahan, “Kant, Naturphilosophie and Ørsted’s Discovery…, pp. 303–304: “Kant held that pure reason alone is unable to produce a science of nature, and that therefore careful empirical investigations are necessary. In this last respect Kant differed sharply from certain other ‘dealers in the a priori’ who assumed that empirical research was unnecessary, because pure reason was capable of arriving at a complete knowledge of nature—Schelling, in particular, seemed to suppose this. Ørsted’s attitude was much closer to Kant’s than to Schelling’s. Like Kant, Ørsted was both more knowledgeable about and more respectful of the actual achievements of physical scientists than were the majority of Naturphilosophen, Schelling included.…The Naturphilosophen were mainly impressed by how much physical knowledge can be arrived at by the use of pure reason. Kant, and Ørsted, tended to focus on how little could be achieved by pure reason alone, and thus regarded experimental investigations more highly.” Christensen (1995, 184–185) makes a point of endorsing Shanahan’s conclusion about the relative importance of Kant’s and Schelling’s influence on Ørsted.

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

151

1799 dissertations on philosophy of nature crafted under the explicit inspiration of Kant’s Metaphysical Foundations—where, in particular, Ørsted already argues for an even more aprioristic deduction on the speculative side than does Kant himself.44 Moreover, the importance of such a fully a priori speculative deduction—now modelled explicitly on Schelling’s Naturphilosophie—became evermore important to Ørsted as his intellectual development progressed. In a paper on “What is Chemistry?” from 1805, for example, Ørsted contrasts “speculative” and “empirical” approaches to this science, where “[t]he latter merely comes up with scattered objects which invite reflection, and from this an arrangement into coherent elements emerges; the former seeks the first principle of everything, sees what constructions must result from it, and prefers to offer the most fundamental construction of science as its definition [of the science].” Thus, “[i]t is not at all necessary for chemistry to be only experimental”; on the contrary, “a universal construction is necessary to complete science, and it stands to reason that such a construction cannot be given through experience but can only be expected from speculation.” As the context makes perfectly clear, the universal speculative construction envisioned here is precisely that originally attempted by Schelling.45

44

45

See Jelved et al., Selected Scientific Works of Hans Christian Ørsted… pp. 46–47), Ørsted, Naturvidenskabelige Skrifter…, vol. 1, 35–36): “In order to achieve completeness in our knowledge of nature, we must start from two extremes, from experience and from the intellect itself. The former method is regressive, beginning with composite facts and resolving these until it arrives at the most simple; the latter is progressive and thus begins with the simplest and progresses towards the most composite. Consequently, the former method must conclude with natural laws which it has abstracted from experience, while the latter must begin with principles, and gradually, as it develops more and more, it also becomes ever more detailed.…When the empiricist in his regression towards the general laws of nature meets the metaphysician in his progression, science will reach its perfection.” Kant, according to Ørsted, has taken the necessary first steps along the path of such a metaphysics; yet, as Ørsted points out at the end of this dissertation, Kant did not fully satisfy the demands of a priori metaphysics in so far as he began with an explicitly empirical concept (the empirical concept of matter). See Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, p. 76, Ørsted, Naturvidenskabelige Skrifter…, vol. 1, 76): “By taking empirical concepts as a basis and inferring the natural laws from them, one imparts to the natural laws thus proved only hypothetical validity instead of the rigorous generality which they should have. According to critical philosophy, all natural laws ought to be deduced from the nature of our cognition, which Kant has developed so excellently in his Critique of Pure Reason, and I believe I have proved that this can be done by deducing them all a priori and taking only what that book has proved for my basis. Therefore I did not hesitate to deviate from Kant’s letter in order to follow the spirit of critical philosophy.” Shanahan, “Kant, Naturphilosophie, and Ørsted’s Discovery…, p. 297, is particularly misleading, therefore, when he (unfavorably) contrasts Schelling’s purely aprioristic ideal with Kant’s reliance on “the empirically given concept of matter” in his Metaphysical Foundations—for it is precisely this latter that is explicitly rejected by Ørsted. See Jelved et al., Scientific Works of Hans Christian Ørsted.., pp. 98–99), Ørsted, Naturvidenskabelige Skrifter…, vol. 3, pp.114–115. On the preceding page Ørsted outlines Schelling’s further articulation of the fundamental forces of attraction and repulsion in general required by the science of chemistry in accordance with the three dimensions of space: magnetism acting in a line, electrical effects on surfaces, and chemical effects (in this case heat) acting “equally freely in all directions in a body” (compare note above). On the succeeding page Ørsted concludes his discussion of a projected complete science of nature, in which experiment and speculation are perfectly balanced, with the words Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, 199) (Ørsted, Naturvidenskabelige Skrifter… vol. 3, p. 116): “It is well-known that Schelling, through speculation, has produced an attempt which, as such, is of incalculable value, but the combined efforts of a great number of blessed geniuses are probably required for the accomplishment of this task.”

152

M. FRIEDMAN

Perhaps Ørsted’s most striking and explicit statement of this dual methodology occurs in his 1811 “First Introduction to the General Doctrine of Nature [Naturlaere]”: In our knowledge of nature we distinguish between something which comes more immediately from reason and something else which rather has its origin in the senses. The two are in the most intimate connection with each other. It is the essence of man to present reason in an organic body, not merely in one particular form, but in its selfcontemplating [selvbeskuende] totality. His sensuous nature, in the most exact sense, can only be regarded as the embodiment of this reason. Therefore, the external sense organs already receive impressions in a manner which is in the most perfect harmony with it, and an unconscious reason in the internal sense impresses its own stamp even more markedly on these various abilities. Imperceptibly, they thus approach the conscious reason which organizes and combines everything into even higher units which, step by step, are finally transformed into the remarkable internal harmony of the independent reason. Thus, the science of experience (empirical science) comes into existence. Reason, on its side, is similar to the internal foundation and essence of nature. In a way, it contains the seeds of the entire world and must develop them through its necessary self-contemplation [Selvbeskuelse]. Consequently, it starts from the highest to which our spirit can ascend, from the essence of beings, the origin of everything. In itself, as a sign of this, it seeks out the various primary directions and through them the origin of the essential fundamental forms in the eternal unity. In its own laws it sees those of nature, in the variety of its own forms that of the world, and thus it develops and creates from itself the great All. In this way arises speculative natural science, which is also called philosophy of nature [Naturphilosophien].

After remarking that “[o]n the empirical path, we are stopped by the enormous profusion of objects which our senses offer, in which, however, there is no completeness,” Ørsted then concludes: [S]peculative natural science seems to lead us more immediately to our goal, but here we would do well to bear in mind that the reason which reveals itself in nature is infinite while ours, which must discover it there, is limited, trapped in finitudes. In innumerable sparks, reason spreads through mankind. Although a reflection of the whole in every individual, it has in each its distinctive direction which prevents it from spreading its light equally clearly and fully in all directions. Only recently shaped in its present form, speculative natural science will only approach significant perfection through the combined efforts of many thinkers.

And the context again makes it clear that the speculative natural science which has been “only recently shaped in its present form” is Schelling’s.46 46

See Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, pp. 288–289; translation emended), Ørsted, Naturvidenskabelige Skrifter…, vol. 3, pp. 162–163. In the following pages Ørsted again puts forward the characteristic view that chemistry is now a branch of general dynamics, where the two original fundamental forces manifest themselves in “electric, galvanic, and magnetic effects” Jelved et al… p. 291; Ørsted, Naturvidenskabelige Skrifter… vol. 3, p. 166); and a comparison of the passage quoted in the text above with the closely related one cited in note strongly suggests that Schelling represents the initial decisive contribution to speculative physics which now needs to “approach significant perfection through the combined efforts of many thinkers.” This is confirmed near the end of the 1811 paper where Ørsted describes “[t]he progress of philosophy in the eighteenth century” (Jelved et al…, p. 305; (Ørsted, Naturvidenskabelige Skrifter …, vol. 3, p. 185: “The perspicacity of Immanuel Kant liberated it from the atomistic system, which, though of speculative origin, was made the basis of experimental physics. F. W. J. Schelling created a new philosophy of nature [Naturphilosophie], the study of which must be important to the empirical student of nature and must both inspire many new ideas in him and also prompt him to re-examination of much that was previously considered unquestionable.” (The idea that Kant liberated natural philosophy from atomism is one of the main thrusts of Ørsted’s 1799 dissertations on the subject.)

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

153

What is most striking about this particular passage is that Ørsted’s methodological ideal of a dual parallel development of both speculative and empirical approaches is grounded, in turn, in a grand metaphysical vision of rational mind and external nature as two complementary aspects of a single total organic development. It is precisely because sensible nature, on the one side, is a realization or externalization of an infinite rational spirit, whereas rational mentality, on the other side, is the evolutionary culmination of an infinite dialectical process by which sensible nature itself continually unfolds, that there is, for Ørsted, a necessary harmony between speculative and experimental inquiry. Since our external senses must be regarded as the “embodiment” of our reason, and because sensible nature itself is the dialectical externalization of an infinite reason (to which our finite reason necessarily approximates), empirical natural science and speculative philosophy of nature must ultimately coincide. Ørsted’s dual methodology is itself grounded, in this sense, in the primary thesis of Schelling’s metaphysics—in the claim, in Schelling’s terms, that transcendental philosophy and Naturphilosophie, spirit within us and nature outside us, must ultimately coincide.47 But this now leads us back to Gower’s question: in what precise sense is this grand—and, from our present point of view, rather wild—metaphysical vision actually implicated in Ørsted’s concrete scientific work? We have already made the essential point in our earlier discussion of Ørsted’s great discovery. For, as we have said (see the paragraph to which Gower’s note is appended), at the center of this metaphysical vision stood the new electrochemistry. It was this new chemistry, interpreted, along with Schelling, as a further articulation and elaboration of general dynamics, that provided the crucial dialectical middle term between matter as conceived in Kant’s original dynamical theory (as a balance or equilibrium of the two fundamental forces) and matter as conceived in Naturphilosophie—as an inexhaustible source of rational life. And it was Schelling’s particular way of conceiving the new chemistry as such an extension of general dynamics—as a triadic dialectical “second potency” of the original dynamical process embracing magnetic, electrical, and chemical or galvanic phenomena—which provided Ørsted with his model for the specific kind of unity of the fundamental forces

47

Ørsted’s dual methodology is thereby grounded, from his own perspective, in Schelling’s definitive overcoming of all skepticism: see note above. To see even more explicitly the connection between this point and the grand naturphilosophisch vision expressed in the passage to which note above is appended, compare the following passage from a discussion of the history of chemistry in 1807 (Jelved et al., Selected Scientific Works…, pp. 259–2560; translation emended); Ørsted, Naturvidenskabelige Skrifter…, vol. 1, pp. 342–343): “Not just chemistry but all human knowledge has always, although with varying clarity, intervened in the essence of things. This has always developed through a continually renewed struggle, which, however, has resolved itself in perfect harmony. And it is not just science, not just human nature, it is all of nature that develops according to these laws.…In short, the development of the earth is precisely like that of the human spirit. This harmony between nature and spirit is hardly accidental. The further we progress, the more perfect you will find it, and you all the more readily agree with me in assuming that both natures are shoots from a common root.…We have glimpsed a higher physics, where the development of science itself, along with all its apparent contradictions, itself belongs to the doctrine of nature. It shows us that everything in the great whole has grown from a common root and will develop into a common life.”

154

M. FRIEDMAN

of nature that guided all of his experimental investigations, including, above all, his discovery of electromagnetism.48 Indeed, in the absence of Schelling’s triadic model, there simply was very little reason, either theoretically or experimentally, to expect the relationship Ørsted actually discovered in the first place; and this, no doubt, is why Ørsted’s discovery was initially received by the scientific community as such a surprise.49 It is precisely here, therefore, that Schelling’s bold metaphysical speculations were fruitfully intertwined with the very latest results of empirical scientific research. Indeed, Schelling’s speculative physics was intimately engaged with the new electrochemistry at its cutting edge, since, as we have seen, Schelling had become convinced of the fundamentally electrical nature of chemical forces already in the first 1797 edition of his Ideas for a Philosophy of Nature, and thus prior to the invention of the Voltaic pile and the consequent discovery of the decomposition of water by an electric current. When Schelling learned of the electrolytic decomposition of water through the work of Ritter, he was then perfectly positioned to seize upon this result and triumphantly announce a new dynamical synthesis of magnetic, electrical, and chemical or galvanic forces in the second 1803 edition.50 So it is no wonder, in particular, that working experimental investigators at the forefront of the new electrochemistry—men like Ritter and Ørsted

48

49

50

See notes above. The question of the specific kind of unity envisaged here is crucial, for, as we have emphasized, it is precisely this that distinguishes Ørsted, Ritter, and Schelling from Kant’s original dynamical theory, on the one hand, and from other conceptions of the unity of nature, on the other: see above, together with the paragraphs to which they are appended. The need for considering the precise kind of unity involved has been especially well emphasized by Kenneth Caneva in a very rich and sophisticated discussion of the issue. See, for example: “One of [Naturphilosophie’s] most important characteristics, rightly emphasized by virtually all commentators, is the belief in the unity, the interconnectedness, of the phenomena of nature. Having said that, however, one must immediately recognize not only its vagueness, but also the fact that there were other, very different routes to the same end—in particular, via various aether theories. When one explores the terms in which that unity was to be achieved, one gets closer to notions especially characteristic of Naturphilosophie—here, in particular, the concepts of Steigerung and Stufenfolge and the elaborate tripartite schemas that embraced broad classes of phenomena in hierarchically situated analogical relationships.” Caneva then emphasizes, in my view perfectly correctly, the importance of such “triadic periodicity”—as explicitly taken from Schelling—in the work of both Ritter and Ørsted. See Caneva, “Physics and Naturphilosophie: A Reconnaissance,” History of Science 35 (1997), pp. 35–106. Of course the obvious similarities and analogies between electrical and magnetic attractions and repulsions were well known since antiquity, and it was also well known that static electricity produces magnetic effects (e.g. in thunderstorms). The difficulty, however, as Ørsted himself points out, is that “electrified and magnetized bodies [appeared to] have no attractive or repulsive effect on one another as a consequence of their condition” (Jelved et al., Selected Scientific Works of Hans Christian Ørsted…, p. 378; Ørsted, Naturvidenskabelige Skrifter…, vol. 2, p. 146. Moreover, there was as yet very little reason for anticipating magnetic effects from chemical or galvanic phenomena. To be sure, Ritter had done some experimental work in magnetochemistry—but, aside from being rather shaky, this work was guided by the very same triadic schema (i.e. Schelling’s) as was Ørsted’s. As Ørsted himself explains (notes above), it appears that he arrived at his decisive experiment by beginning from Schelling’s triadic schema and then adding the idea that electrical forces are much more “bound” (i.e. less manifest as static electricity or charge) in a galvanic circuit than they are in a static distribution of charge. From this point of view, current electricity is much closer, as it were, to magnetism, and so it is precisely here that we should seek the desired “attractive or repulsive effect.” See notes above. The Supplement to Book I, ch. 3 of Ideas mentioned in note concerns “the History of the Decomposition of Water,” and it begins with the words (1989, p. 93) (1848, p. 119): “A more non-sensical undertaking could hardly be imagined than to attempt to outline a universal

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

155

himself—were also able to take positive philosophical inspiration from Schelling’s metaphysical speculations. It is in this precise sense that we here have an especially good example of what I have called a timely confluence of empirically grounded scientific problems and speculatively generated philosophical ideas. More specifically, the timeliness of Schelling’s particular philosophical intervention into the ongoing practice of empirical scientific research can now be further elucidated as follows. Kant had already crafted his original dynamical theory of matter in intimate connection with the most fundamental principles of Newtonian physics (see notes above), and Kant was increasingly conscious, at the same time, of the need to extend his constitutive grounding of Newtonian physics so as to embrace the new discoveries in matter theory and chemistry made at the end of the 18th century (see note above, together with the paragraph to which it is appended). In just this connection, however, Kant faced deep philosophical problems centering around his sharp distinction between constitutive principles and merely regulative principles, eventually leading to a final attempt, in the Opus postumum, to articulate a radically new type of a priori construction—a spacefilling aetherial medium or caloric fluid—combining both regulative and constitutive aspects in a single unitary representation (see note above, together with the paragraph to which it is appended). Here, as we have seen, he was able, with some difficulty, to provide a kind of constitutive grounding for Lavoisier’s recently articulated chemical system, but even this final effort fell considerably short of the new electrochemistry which became centrally important at the turn of the century. The timeliness, and consequent fruitfulness, of Schelling’s particular intervention therefore consists in the way in which he responded to the deep philosophical problems created by Kant’s sharp distinction between constitutive principles and regulative principles, on the one hand, while simultaneously engaging with the newest developments at the forefront of electrochemistry, on the other. Kant’s own attempts further to develop his system so as to resolve both the philosophical problems in question and to accommodate recent scientific results had led him, as we have seen, right up to the verge of the radically new conceptual situation faced by Schelling, Ritter, and Ørsted. Nevertheless, the decisive step of conceiving magnetic, electrical, and chemical or galvanic effects in terms of a triadic continuation of a dialectical process beginning with the fundamental forces of attraction

theory of nature from particular experiments; nevertheless, the whole of French chemistry is nothing but such an attempt—but it could hardly [be expected], as well, that the overwhelming value of higher prospects, directed upon the whole, over those based on particularities, would have been finally so admirably confirmed, as in precisely the history of this doctrine, especially in that part of it that concerns the nature of water.” Thus, although it is certainly true that Schelling was no empiricist (and there is also no doubt that Schelling’s speculative flights often led him astray), Shanahan is completely off the mark when he suggests (note above) that, for Schelling, “empirical research was unnecessary” and insinuates that Schelling was neither “knowledgeable about” nor “respectful of ” concrete scientific work. Schelling’s attitude towards speculation and experiment was not so different from Ørsted’s: the former is necessary to organize the results of the latter into a coherent and scientific unified totality on the basis of “higher prospects, directed upon the whole.” For discussion of Schelling’s similarly dual methodology—whereby a priori deduction must be verified by the empirical facts, which it serves, in turn, to systematize—see Beiser, German Idealism…, § IV.4.6).

156

M. FRIEDMAN

and repulsion in general is one that Kant did not and could not take—and it is just this crucial step that was then left to Schelling alone.51 It is clear, then, that the timeliness, and consequent fruitfulness, of a philosophical intervention into ongoing scientific research does not depend on the validity, by current standards, of the philosophical ideas in question—nor does it require that the theoretical schemata suggested by these philosophical ideas should themselves be correct according to subsequent scientific theorizing. In the present case, in particular, neither Schelling’s bold metaphysical vision nor the triadic developmental dynamical model of magnetic, electrical, and chemical or galvanic phenomena managed to survive into the second half of the 19th century, let alone into the twentieth.52 Yet Schelling’s speculative physics did in fact fruitfully guide the experimental and theoretical work of the principal founders of electrochemistry and electromagnetism at the beginning of the 19th century.53 And it was able to do this, in spite of the shortcoming we might perceive in it today, precisely because the new empirical situation at the turn of the century demanded the exploration of non-Newtonian physical ideas, while, at the same time, the Kantian philosophical system, at the very end of the 18th century, had pushed such Newtonian ideas to their outermost philosophical limits. A new philosophy of nature, like Schelling’s,

51

52

53

See note above, together with the paragraph to which it is appended. Thus, once again, there was simply no way—as both Shanahan and Christensen appear to assume (note)—that Kant’s original dynamical theory of the fundamental forces of attraction and repulsion in general could, by itself, have suggested the specific kind of unity of the various forces and powers of nature that was central, in particular, to the new electrochemistry. Here Pearce Williams hits the nail directly on the head when he writes: “Kant’s original attractive and repulsive forces were adopted and transformed by Schelling. What Schelling added was the substitution of development for equilibrium between conflicting forces. Thus when equal and opposing forces met the result for Schelling was not stable equilibrium, but development into a higher conflict. It was in Schelling that the famous triad of thesis, antithesis, and synthesis first appeared in the guise of dynamic, conflicting forces. The world is thus the scene of ceaseless activity and dynamic progression.” Williams, “Kant, Naturphilosophie and Scientific Method…, p. 15. It is largely for these reasons, it seems, that Gower, in particular, remains skeptical of the scientific relevance of Schelling’s ideas (see note above). At one place, for example, he refers to Schelling’s “bizarre theory that nature and self are ultimately identical” (p. 313); at another he asserts that “there is little of philosophical value to be gained by an examination of [the] arguments and theories [of Schellingean metaphysics]” (p. 352). He also complains (p. 349) that “[Ørsted’s] dynamical theories of physical action, with their emphasis upon the interaction of polar forces, contributed little to the creation of a conceptual framework with which the energy conservation principles could emerge and be understood.” Nevertheless, as I have suggested (note above), Gower’s discussion is generally quite balanced and helpful, and, as a whole, it conveys a remarkably good sense of how Schelling’s ideas concerning “polarity”—or what Caneva calls “triadic periodicity” (note above)—in fact fruitfully guided the work of both Ritter and Ørsted. Aside from the work of Ritter and Ørsted, as we pointed out above (note), this also includes such seminal figures as Davy and Faraday. And, although later figures, such as Helmholtz, for example, self-consciously rejected the pan-organic conception of nature originating in Naturphilosophie (see note above, together with the paragraph to which it is appended), the question of the relationship of the new electrochemistry and electrodynamics to the life sciences—in Helmholtz’s case, to sensory and motor physiology—continued to be a central driving force. Indeed, this is still true of Maxwell’s electrodynamical theorizing—which, as in the case of Helmholtz, continued to draw on 19th-century physiology: see Cat, “On Understanding: Maxwell on the Methods of Illustration and Scientific Metaphor,” Studies in History and Philosophy of Modern Physics 32 (2001), pp. 395–441. Only with the triumph of such slogans as “Maxwell’s theory is Maxwell’s equations” at the beginning of the 20th century did this originally biological problematic finally completely disappear.

KANT—NATURPHILOSOPHIE—ELECTROMAGNETISM

157

which, as we have said, intelligently responds simultaneously to both the deep tensions emerging in Kant’s philosophical system and the new empirical results, was therefore, in this specific historical context, precisely what was then needed.54 Ørsted’s spectacular discovery of electromagnetism provides us with perhaps our single most important confirmation. Stanford University

BIBLIOGRAPHY Beiser, Frederick (2002). German Idealism. Cambridge, MA: Harvard University Press. Caneva, Kenneth (1997). “Physics and Naturphilosophie: A Reconnaissance.” History of Science 35: 35–106. Cassirer, Ernst (1920). Das Erkenntnisproblem in der Philosophie und Wissenschaft der neueren Zeit. Dritter Band: Die Nachkantischen Systeme. Berlin: Bruno Cassirer. Cassirer, Ernst (1950). The Problem of Knowledge: Philosophy, Science, and History since Hegel. Translated by W. Woglom and C. Hendel. New Haven, CT: Yale University Press. Cat, Jordi (2001). “On Understanding: Maxwell on the Methods of Illustration and Scientific Metaphor.” Studies in History and Philosophy of Modern Physics 32: 395–441. Christensen, Dan (1995). “The Ørsted-Ritter Partnership and the Birth of Romantic Natural Philosophy.” Annals of Science 52: 153–185. Förster, Eckart (2000). Kant’s Final Synthesis. Cambridge, MA: Harvard University Press. Franks, Paul (forthcoming). “What should Kantians learn from Maimon’s Skepticism?” In G. Freudenthal, ed. The Philosophy of Salomon Maimon and its Place in the Enlightenment. Dordrecht, The Netherlands: Kluwer. Friedman, Michael (1992). Kant and the Exact Sciences. Cambridge, MA: Harvard University Press. Friedman, Michael (2001). Dynamics of Reason. Stanford, CA: CSLI Publications. Gower, Barry (1973). “Speculation in Physics: The History and Practice of Naturphilosophie.” Studies in History and Philosophy of Science 3: 301–56. Jelved, Karen, Andrew Jackson, and Ole Knuden, eds. (1998). Selected Scientific Works of Hans Christian Ørsted. Princeton, NJ: Princeton University Press. Kant, Immanuel (2002). Metaphysical Foundations of Natural Science. Translated by M. Friedman. In H. Allison and P. Heath, eds. Immanuel Kant: Theoretical Philosophy after 1781. Cambridge: Cambridge University Press. Kuhn, Thomas (1959). “Energy Conservation as an Example of Simultaneous Discovery.” In M. Clagett, ed. Critical Problems in the History of Science. Madison, WI: University of Wisconsin Press. Newton, Isaac (1931). Opticks. London: G. Bell and Sons. Ørsted, Hans Christian (1920). Naturvidenskabelige Skrifter (Scientific Papers). 3 vols. Copenhagen: Royal Danish Society of Sciences. Royce, Josiah (1919). Lectures on Modern Idealism. New Haven, CT: Yale University Press. Schelling, Friedrich Wilhelm (1848). Friedrich Wilhelm Josef von Schellings Sämmtliche Werke. Zweiter Band. Stuttgart und Augsberg: J. G. Cotta’scher Verlag. Schelling, Friedrich Wilhelm (1988). Ideas for a Philosophy of Nature. Translated by E. Harris and P. Heath. Cambridge: Cambridge University Press.

54

The crux of the matter, of course, is that Schelling’s Naturphilosophie was in fact an intelligent, perceptive, and appropriate response to both the tensions in Kant’s system and the new empirical results. Viewed in this particular context, I believe, there is indeed much of “philosophical value” to be gained by studying it. For a more general attempt—focussing on quite different examples—to elucidate the crucial interaction between philosophical ideas and scientific problems along these lines, see Friedman, Dynamics of Reason (Stanford, CSLI Publications, 2001).

158

M. FRIEDMAN

Shanahan, Timothy (1989). “Kant, Naturphilosophie, and Ørsted’s Discovery of Electromagnetism: A Reassessment.” Studies in History and Philosophy of Science 20: 287–305. Stauffer, Robert (1957). “Speculation and Experiment in the Background of Ørsted’s Discovery of Electromagnetism.” Isis 48: 33–50. Williams, L. Pearce (1965). Michael Faraday: A Biography. New York: Basic Books. Williams, L. Pearce (1966). The Origins of Field Theory. New York: Random House. Williams, L. Pearce (1973). “Kant, Naturphilosophie and Scientific Method.” In R. Giere and R. Westfall, eds. Foundations of Scientific Method in the Nineteenth Century. Bloomington, IN: Indiana University Press.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH ERNST P. HAMM1

“This book offered a beautiful interpretation, informed by the character of the recent Naturphilosophie, of the geology of that time, and it was written with the intellect and eloquence that so distinguished Steffens. Its many bold and astute reflections drew much attention and were not without effect on its numerous readers; but now, looking back over almost half a century, we must confess that it did not enrich science with such clear results that they are here worth mentioning.”2 Hans Christian Ørsted’s mixed praise for Contributions to the Internal Natural History of the Earth (1801) grows faint when we consider that this book was the most renowned scientific work of his recently deceased countryman, Henrik Steffens, who was being honoured with a commemorative address. Ørsted took the occasion to warn of the ways in which the empirical natural sciences (Erfahrungsnaturwissenschaften) were undervalued in those heady days at the onset of the 19th century. Steffens was fortunate enough to have had a solid foundation in empirical science before he took to philosophy and so he “was never completely caught in the confusions of those times”; others, unable to meld the stuff of experience with higher philosophical flights, ended up with a “foggy self-satisfaction and a dissatisfaction with all scientific knowledge.”3 That Ørsted would tread so carefully—and this before an audience where his reputation as a natural philosopher was nothing if not secure—in his reflections on the relationship between science and philosophy, a relationship he had always been eager to embrace, gives some indication of the character of mid-19th century science. Schelling, who had long since left behind his Naturphilosophie, took similar precautions in his éloge for Steffens, noting that natural science could more or less go about its business without attending to philosophy.4 Other contemporaries, as is well known,

1 2

3 4

I wish to acknowledge the support of the Atkinson Faculty Committee on Research. Hans Christian Ørsted, “Henrik Steffens: Gedächtnißrede, gehalten in der dänischen Gesellschaft der Wissenschaften am 6. März 1846,” in Kleinere Schriften von Hans Christian Ørsted, translated by K. L. Kannegiesser, 2 vols. (Leipzig: C. B. Borch, 1855), vol. i, pp. 107–120, at p. 111. All translations are my own unless otherwise indicated. Ibid. pp. 113, 116. “One would not so readily have said then what can be heard almost daily now: that natural science goes about its business better when it keeps a distance from all philosophy; a statement that is as true as the claim that cookery does not proceed best for the person who wants to base everything on chemical principles.” Friedrich Wilhelm Joseph von Schelling, “Aus einem öffentlichen Vortrag zu H. Steffens Andenken, gehalten am 24. April 1845,” in H. Steffens, Nachgelassene Schriften, with

159 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 159–175. © 2007 Springer.

160

E. P. HAMM

were less moderate in their assessment and all too eager to heap opprobrium on Naturphilosophie. The éloge, a genre heavily laden with the burden of occasion, needs to be handled with considerable historical care;5 yet the remarks of Ørsted came to typify for more than a century the historiography of Steffens and his fellow travelers: dazzling speculation blinded Naturphilosophen to the importance of empirical science. This was but a part of an even larger picture that portrayed Romanticism as hostile to the achievements of the Enlightenment, especially its science. The Romantic legacy, such as it was, lay in the Geisteswissenschaften and the ideals of Bildung. This is good fodder for those who would retail the notion of two cultures,6 but it has its shortcomings as history, in particular it does not help us understand what it was about science that engaged the minds of the some of the most outstanding figures of the Romantic era. Over the last five decades this picture of the Romantic legacy has undergone considerable revision and we now know that Naturphilosophie and other manifestations of Romanticism were tied to a number of developments in the natural sciences7 and, more fundamentally, to the transformation of 18th-century practices of natural history and natural philosophy into new disciplines of physics and biology.8

5

6

7

8

a preface by F. W. J. v. Schelling (Berlin: E. H. Schroeder, 1846), pp. iii–lxiii, at p. vi. The bulk of Schelling’s address focuses on church, state, and the character of German Protestantism. For example, see Dorinda Outram, “The Language of Natural Power: The ‘éloges’ of Georges Cuvier and the Public Language of Nineteenth-Century Science,” History of Science 16 (1978): pp. 153–178. See the helpful discussion by Stefan Collini in his “Introduction” to C. P. Snow, The Two Cultures (Cambridge: Cambridge University Press, 1993), pp. vii–lxxi, at. pp. ix–xi. The literature on science and Romanticism is considerable. Ørsted is a preeminent case; see, for example: Robert C. Stauffer, “Speculation and Experiment in the Background of Ørsted’s Discovery of Electromagnetism,” Isis 48 (1957): pp. 33–50; Thomas Kuhn, “Energy Conservation as an Example of Simultaneous Discovery,” [1959] in The Essential Tension: Selected Studies in Scientific Tradition and Change (Chicago, IL: University of Chicago Press, 1977), pp. 66–104; Dan Ch. Christensen, “The Ørsted-Ritter Partnership and the Birth of Romantic Natural Philosophy,” Annals of Science 52 (1995): pp. 153–185; Kenneth L. Caneva. “Colding, Ørsted, and the Meaning of Force,” Historical Studies in the Physical and Biological Sciences 27 (1997): pp. 1–138; Andrew D. Wilson, “Introduction” to Selected Scientific Works of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen, with an introduction by Andrew D. Wilson (Princeton, NJ: Princeton University Press, 1998), pp. xv–xl; and Michael Friedman “KantNaturphilosophie-Electromagnetism,” in this collection. See also Kenneth L. Caneva, “Physics and Naturphilosophie: A Reconnaissance,” History of Science 35 (1997): pp. 35–106; Barry Gower, “Speculation in Physics: The History and Practice of Naturphilosophie,” Studies in History and Philosophy of Science 3 (1972–1973): pp. 301–356; Dietrich von Engelhardt, Historisches Bewußtsein in der Naturwissenschaft von der Aufklärung bis zum Positivismus (Freiburg and München: Albers, 1979); David M. Knight, “The Physical Sciences and the Romantic Movement,” History of Science 9 (1975): pp. 54–75; Timothy Lenoir, The Strategy of Life: Teleology and Mechanics in NineteenthCentury German Biology [1982] (Chicago, IL: University of Chicago Press, 1989); Trevor H. Levere, Poetry Realized in Nature: Samuel Taylor Coleridge and Early Nineteenth-Century Science (Cambridge: Cambridge University Press, 1981); and Robert J. Richards, The Romantic Conception of Life: Science and Philosophy in the Age of Goethe (Chicago, IL: University of Chicago Press, 2002); and the useful collection of papers in Andrew Cunningham and Nicholas Jardine, Romanticism and the Sciences (Cambridge: Cambridge University Press, 1990). On changing disciplines see: Rudold Stichweh, Zur Entstehung des modernen Systems wissenschaftlicher Disziplinen: Physik in Deutschland, 1740–1890 (Frankfurt a. M.: Suhrkamp, 1984); William Clark, “German Physics Textbooks in the Goethezeit,” parts 1 and 2, History of Science 35 (1997): pp. 219–239, 295–363; and Nicholas Jardine, “The Significance of Schelling’s ‘Epoch of a Wholly New Natural History’: An Essay on the Realization of Questions,” in Metaphysics and Philosophy of Science in the Seventeenth and Eighteenth Centuries: Essays in Honour of Gerd

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

161

Chemistry was a practice with a very long tradition, but as Michael Friedman has shown in his contribution to this collection, it was radically rethought by Schelling, who saw chemical, galvanic, magnetic, and electrical phenomena as having a common source in an original dynamic process. Ørsted shared this view, which turned out to be the naturphilosophische idea behind his discovery of electromagnetism. Geology, another discipline that came into its own in the decades surrounding 1800, and its interactions with Naturphilosophie has been the subject of relatively little attention.9 This despite the fact that geology was the natural science in which historical thinking and notions of development over time first came to prominence—a point Schelling alluded to in his éloge of Steffens—and that the idea of development pervaded Naturphilosophie.10 The case of Steffens is instructive in this regard, for it shows that Naturphilosophie, far from being hostile to empirical science, went hand in hand with one of its outstanding examples, the geognosy of Abraham Gottlob Werner. For Steffens, geology was part of a much

9

10

Buchdahl, ed. R. S. Woolhouse, The University of Western Ontario Series in the Philosophy of Science; 43 (Dordrecht The Netherlands; Boston, MA: Kluwer, 1988), pp. 327–350. Some exceptions that prove the rule are: Bernhard Fritscher, “A. G. Werner (1749–1817) als Lehrer der deutschen Naturphilosophie—Zum Werk von Henrik Steffens (1773–1845),” Zeitschrift für geologische Wissenschaften 21 (1993): pp. 495–502; idem, “Bemerkungen zu einer historischen Epistemologie der romantisch-idealistischen Erdwissenschaften am Beispiel der ‘Geosophie’ Lorenz Okens,” Zeitschrift für geologische Wissenschaften 27 (1999): pp. 61–69; Fergus Henderson, “Practical Knowledge in Romantic Ordering of Nature: Werner’s Geognostical Method, Humboldt’s ‘Pasigraphie,’ Novalis’s ‘Instrumentalsprache,’ and the Encyclopaedism of the Individual,” in Toward a History of Mineralogy, Petrology, and Geochemistry, ed. Bernhard Fritscher and Fergus Henderson, Algorismus; 23 (München: Instititut für Geschichte der Naturwissenschaften, 1998), pp. 131–146; Gottfried Hofbauer, “Die sinnliche Naturgeschichte des Abraham Gottlob Werner. An der Grenze zwischen Empirismus und romantischer Naturphilosophie,” Zeitschrift für geologische Wissenschaften 21 (1993): pp. 545–558. The general connections between romanticism, geology, mining, and the subterranean world have been the subject of more study: Roger Cardinal, “Werner, Novalis and the Signature of Stones,” in Deutung und Bedeutung: Studies in German and Comparative Literature presented to Karl Werner Maurer, ed. Brigitte Schludermann et al. (The Hague: Mouton, 1973), pp. 118–133; Michael Dettelbach, “Humboldtian Science,” in Cultures of Natural History, ed. Nicholas Jardine et al. (Cambridge: Cambridge University Press, 1996), pp. 287–304; Josef Dürler, Die Bedeutung des Bergbaus bei Goethe und in der deutschen Romantik (Frauenfeld-Leipzig: Huber, 1936); E. P. Hamm, “Unpacking Goethe’s Collections: the Public and the Private in Natural Historical Collecting,” British Journal for the History of Science 34 (2001): pp. 275–300; Alexander M. Ospovat, “Romanticism and Geology: Five Students of A. G. Werner,” Eighteenth Century Life 7 (1982): pp. 105–117; Nicholas A. Rupke, “Caves, Fossils and the History of the Earth,” in Romanticism and the Sciences (cited n. 7), pp. 241–259; Michael Shortland, “Darkness Visible: Underground Culture in the Golden Age of Geology,” History of Science 32 (1994): pp. 1–61; John Wyatt, Wordsworth and the Geologists (Cambridge: Cambridge University Press, 1995); Theodore Ziolkowski, “The Mine: Image of the Soul,” in German Romanticism and Its Institutions (Princeton, NJ: Princeton University Press, 1990), pp. 18–63. The “fundamental point” (Grundanschauung) that drew Schelling and Steffens to one another from the time of their first acquaintance was that Steffens “had won for the unfathomable history of the earth a whole series of epochs, in which one covers another, one becomes the foundation for the next, and not without undergoing change in the process.” Schelling, “Aus einem öffentlichen Vortrag zu H. Steffens Andenken” (cited n.4), p. v. On historical thinking in the natural sciences see Engelhardt, Historisches Bewußtsein (cited n. 7). For geology as an historical science see: Stephen J. Gould, Time’s Arrow, Time’s Cycle: Myth and Metaphor in the Discovery of Geological Time (Cambridge, MA: Harvard University Press, 1987); David R. Oldroyd, “Historicism and the Rise of Historical Geology,” parts 1 and 2, History of Science 17 (1979): pp. 192–217, 227–257; Rhoda Rappaport, When Geologists Were Historians, 1665–1750 (Ithaca, NY: Cornell University Press, 1997); Martin J. S. Rudwick, Georges Cuvier, Fossil Bones and Geological Catastrophes: New Translations and Interpretations of the Primary Texts (Chicago, IL: University of Chicago Press, 1997).

162

E. P. HAMM

larger story that tied the development of the earth to human Bildung, a story that could easily be imagined to be historical. However, it turns out that the affinities that bound nature to humans in this vision, a vision that was to some extent shared by Steffens and Ørsted, was not primarily historical, it was chemical. Steffens spent most of his adult life in the German-speaking world, but his love for natural history began already in his Norwegian adolescence when, according to his autobiography, he was fascinated with minerals, fossils, and the writings of Buffon, and was further developed at the University of Copenhagen, where he studied the natural sciences, above all mineralogy. Chemistry was also among his interests—he was taken with the clarity and perspicuity of Lavoisier’s chemisty and one of his first publications is concerned with the calcinations of metals and dates from his studies at Copenhagen.11 He pursued his mineralogical studies at the University of Kiel, where he succeeded brilliantly in his examinations and wrote a dissertation published in 1797 under the title On Mineralogy and the Study of Mineralogy.12 Steffens wrote On Mineralogy before he had read Schelling. It is not a work of Naturphilosophie nor can it be characterized as speculative, features that make it useful for comparison with Steffens’s thought after he encountered Naturphilosophie. On Mineralogy begins by reviewing the history of mineralogy, a term that in late 18th-century German was broad enough to encompass the study of minerals, fossils, and the structure of the earth’s crust. General theories of the earth, especially the so-called sacred theories of the 17th century, most notably represented by Thomas Burnet, William Whiston, and John Woodward, are dismissed as fantasies. This was very much in keeping with the main current of mineralogical opinion at the close of the 18th century, a time when Buffon’s theories of the earth had come to be seen as the secular counterparts of his theological predecessors: grand, all-encompassing, and empirically thin.13 Such theorizing was dismissed as Geologie, a term associated with unconstrained theorizing by most of Steffens’s contemporaries.14 The empirical description of the earth’s crust was more typically known as Geognosie. After summing up the history of the field, Steffens turned to 11

12

13

14

Biographical details are from H. Steffens, Was ich erlebte: Aus der Erinnerung niedergeschrieben, 10 vols. in 5 ( Stuttgart-Bad Cannstatt: Frommann-Holzboog, 1995–1996, reprint of first edition, Breslau: Max und Komp. 1840–1844), vol. i, pp. 239–242, vol. ii, pp. 192, 211–212. The early publication was H. Steffens, “De fornemste Hypotheser, ved hvis Hjelp man har søgt at forklare Metallernes Forkalkning,” Physicalsk, oeconomisk og medicochirugisk Bibliothek for Danmark og Norge, parts 1 and 2, 1(1794): pp. 42–77, 161–164. H. Steffens, Über Mineralogie und das mineralogische Studium (Altona: Hammerich, 1797), see also idem, Was ich erlebte (cited n.11), vol. iii, pp. 201–202. Georges-Louis Leclerc Comte de Buffon, Histoire naturelle, générale et particulière, vol. 1 (Paris, 1749); idem, Les époques de la nature, ed. Jacques Roger (Paris: Editions du Muséum, 1962), in which Jacques Roger has shown that Buffon deserved to be taken seriously. My concern here is only about how Buffon was received. James Hutton’s Theory of the Earth with Proofs and Illustrations (Edinburgh: Creech, 1795) is an exception to this anti-theoretical trend, which may help explain why it received relatively little attention, at least at first. The late eighteenth-century has been characterized as a time when geology wanted “facts,” not new ideas, Karl Alfred von Zittel Geschichte der Geologie und Paläontologie bis ende des 19. Jahrhunderts, Geschichte der Wissenschaften in Deutschland. Neuere Zeit; 23 (Munich and Leipzig: R. Oldenbourg, 1899), p. 76 Dennis R. Dean, “The word ‘geology’,” Annals of Science 36 (1979): pp. 35–43; see also Rhoda Rappaport, “Borrowed Words: Problems of Vocabulary in Eighteenth-Century Geology,” British Journal for the History of Science, 15 (1982): pp. 27–44.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

163

the work of important contemporaries, the most noted of which was undoubtedly the geognost Abraham Gottlob Werner. Werner’s approach to mineralogy is ably summarized, but in the end it is judged as unsatisfactory because it was guided by the utilitarian aim of improving mining (which is about what is to be expected of a professor at the renowned Freiberg Mining Academy).15 Steffens then proceeded to outline his view of the field. In its broadest sense mineralogy was concerned with the knowledge of inorganic bodies and their relation to each other. He did not deny the interest of historical questions and quoted with approval a remark of the Göttingen physicist and aphorist, Georg Christoph Lichtenberg: “Every landform, the shape of every sand hill and cliff contains the history of the earth written in a natural script.”16 Despite this, Steffens believed that any attempt to address the question of origins would lose itself in fantastic hypotheses.17 There is no reason to believe that this was a position taken out of deference to religious authorities or a literalistic interpretation of the Bible. Few of Steffens’s informed contemporaries doubted the great antiquity of the earth, but they knew very well that evidence for assigning a specific age of the earth, much less its earliest origins, was wanting. Although Steffens denied that mineralogy could ever be called a science in the “strictest sense,” a view doubtless informed by Kant, who said exactly the same thing about chemistry,18 he did think it could be improved by dividing it into three doctrines. The first, Oryktognosie, was concerned with the chemical composition and classification of the solid bodies that make up the mineral kingdom.19 The 18th-century tabular taxonomies of minerals, such as that developed by the Swede Axel Frederik Cronstedt, who divided minerals into four classes: earths, salts, inflammables, and metals,20 were kinds of Oryktognosie. The second doctrine, Geognosie, was concerned with minerals as they naturally occur in rocks and was subdivided into mineralogical geognosy, which focused on the description of rock formations in particular locales and geognosy proper, which treated the overall structure of the earth’s crust.21 This much was straightforward and not especially original; more unusual was the third doctrine, Geologie, which was not speculative but instead focused on the chemical and physical forces that are ever at work shaping the inorganic world.22 Judged by the standards of its time, On Mineralogy and the Study of Mineralogy was undoubtedly a work of empirical science that took a cautious approach to its subject.

15 16 17 18

19 20 21 22

Steffens, Über Mineralogie und das mineralogische Studium (cited n. 12) p. 96. Ibid. pp. 83, 85 cites Georg Lichtenberg, Göttingen Taschen Kalendar, 1778, p. 7. Steffens, Über Mineralogie und das mineralogische Studium (cited n. 12), pp. 68–69, 87–88, 160. Immanuel Kant, Metaphysical Foundations of Natural Science, [1784] transl. with intro. and essay by James Ellington (Indianapolis, IN: Bobbs-Merrill, 1970), pp. 4, 7. Steffens, Über Mineralogie und das mineralogische Studium (cited n. 12), p. 113. Ibid. pp. 52–69. Ibid. p. 154. Ibid. p.158. Geologie treats minerals as a means of a Naturzweck, a “natural purpose” that can only be found in the inorganic world existing as a whole, not in the individual parts. Here Steffens made reference to Kant’s Critique of Judgment. See the helpful discussion of Naturzwecke in biology in Richards, The Romantic Conception (cited n. 7), pp. 65–70.

164

E. P. HAMM

Shortly after completing his dissertation and graduating from Kiel, and almost immediately after the death of his father, Steffens discovered Schelling. It was an important moment, emotionally and intellectually. He recalled that the “Introduction” to Schelling’s Ideas for a Philosophy of Nature lifted his spirits out of mourning and that he read it “with passion.”23 He set about securing a Danish stipend to pursue further study in Germany, though he never spoke of his interest in speculative philosophy to his benefactors, who undoubtedly expected him to study mineralogy at Freiberg.24 He made his way to Jena and quickly developed an excellent working relationship with Schelling, who must have been pleased to have found an enthusiastic collaborator who had earned his credentials as a natural scientist before turning to philosophy.25 The first fruit of this collaboration is evident in the first issue of Schelling’s Journal for Speculative Physics, a substantial amount of which was written by Steffens, who was wholehearted in his enthusiasm for Schelling’s project. In his first contribution to the Journal, a lengthy review and summary of Schelling’s three most recent works,26 Steffens announced that “The highest problem of all natural science would be finding the first law from which all other laws of nature could be derived, or to find nature in its highest simplicity.”27 Such a law would have to have an “inner necessity” and it would be the basis of a true natural science [that would] offer a norm for all the investigations of natural scientists and thus would have to become the mother of all future discoveries.

One will easily see that such an enterprise must lead to the total reform of the present study of nature. In such a science all the divisions of the study of nature into physics, chemistry, physiology, etc., sciences that are divided from one another, would have to fall away, for its purpose would be the unification of all these branches under higher principles.28 Schelling’s project was nothing if not grand and it cannot be done justice in a brief summary, but it is not unjust to say that it was, in a word, dynamic. Mechanical and atomistic explanations were anathema to Naturphilosophie. Kant had argued 23 24 25

26

27 28

Steffens, Was ich erlebte (cited n. 11), vol. iii, pp. 337–339, at p. 338. Ibid. vol. iv, pp. 1, 20; vol. iii, p. 341. Steffens claimed he was the first trained expert in the natural sciences (“Naturforscher von Fach”) to offer unconditional support for Schelling, ibid. vol. iv, p. 76. Though technically true this needs to be qualified with the fact that Johann Wilhelm Ritter, though still a student at Jena and perhaps not unconditional in his support of Schelling, had already made a name for himself with his Beweis, dass ein beständiger Galvanismus den Lebensprocess in dem Thierreich begleite. Nebst neuen Versuchen und Bemerkungen über den Galvanismus (Weimar: im Verlage des Industrie-Comptoirs, 1798). H. Steffens, “Recension der neuern naturphilosophischen Schriften des Herausgebers: F.W. J. Schelling, Von der Weltseele: eine Hypothese der höhern Physik zur Erklärung des allgemeinen Organismus (Hamburg: Perthes, 1798); Erster Etwurf eines Systems der Naturphilosophie. Zum Behuf der Vorlesungen (Jena and Leipzig: Gabler, 1799); Einleitung zum Entwurf eines Systems der Naturphilosophie: oder über den Begriff der speculativen Physik und die innere Organisation eines Systems dieser Wissenschaft (Jena and Leipzig: Gabler, 1799), Zeitschrift für speculative Physik, parts 1 and 2, 1, no. 1 (1800): pp. 1–48, and 1, no. 2 (1800): pp. 88–121. Such reviews cum summaries were typical in German journals of that time. Schelling took editorial liberty and on several occasions inserted footnotes to indicate where Steffens was misinterpreting his work. Ibid. part 1, p. 4–5. Ibid. part 1, p. 5.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

165

that matter was the product of attractive and repulsive forces29 and Schelling, informed by more recent developments in chemistry, developed this idea into what he called dynamics, “the basic science of a theory of Nature.”30 Unlike Kant, Schelling very much saw chemistry as a science, one that followed directly from dynamics. As the product of fundamental forces, matter was essentially dynamic, not inert, and this dynamism was present through the inorganic and organic world. Put another way, this dynamism was nature as productivity (natura naturans), the proper subject of speculative physics; nature as product (natura naturata), was the subject of empical physics.31 Schelling’s project also works in two directions: transcendental philosophy begins by subordinating the real world of nature to the ideal world of reason; Naturphilosophie starts with the real and ascends to the ideal world of reason.32 Nature and mind were intrinsically rational, and the empirical facts of science had of necessity to conform to some rational principle. In a passage that is an accurate gloss of this aspect of Schelling, Steffens wrote: “Nature is originally organic, i.e. productive. That she is so can only be conceived as the conflict of opposed activities.”33 The notion of grounding a comprehensive science of nature on a single unifying principle also had enormous appeal for Ørsted.34 The details of how this fundamental principle worked itself out in nature were matters that needed working out, and this is precisely what Steffens set about doing in “On the Oxidation and Reduction Process of the Earth.”35 This essay begins with the claim that all of the solid bodies of the earth share a universal tendency to crystallization, a tendency that is constantly resisted lest the world become nothing more than an aggregate of crystals. These opposed tendencies are exemplified in granite, the product of the crystallizing tendencies of quartz, mica, and feldspar all in conflict with each other. Under certain conditions, as in caves, fissures or on mountain peaks, these minerals can follow their tendency to crystallize with less resistance and create bodies of crystallized quartz or other minerals, though in such instances there is no granite. Granitic and mineralogical organization more generally is the product of conflicting tendencies, and in the remainder of this brief essay Steffens seeks to explain the structure of the earth in terms of the opposed processes of combustion and reduction. Minerals are divided into two classes: 29 30

31

32 33 34

35

Kant, Metaphysical Foundations (cited n. 18), pp. 40–94. F. W. J. Schelling, Ideas for a Philosophy of Nature: As Introduction to the Study of This Science, transl. Errol E. Harris and Peter Heath, introduction by Robert Stern (Cambridge: Cambridge University Press, 1988), p. 5. This translation is of the 1797 first edition and the 1803 supplements; all references in this paper are from the 1797 sections. See also Friedman, “Kant-NaturphilosophieElectromagnetism” (cited n. 7). Schelling, Einleitung zu dem Entwurf eines Systems der Naturphilosophie. Oder über den Begriff der spekulativen Physik, in Ausgewählte Schriften, 6 vols (Frankfurt a. M.: Suhrkamp, 1985), vol. i, pp. 337–394, at p. 352. Ibid. pp. 340–341. Steffens, “Recension” (cited n. 26), part 1, pp. 36–37. See Wilson, “Introduction,” to Selected Scientific Works (cited n. 7), pp. xxii–xxiii, and pp. xx–xxii for a discussion of the electrical conflict in Ørsted. H. Steffens, “Über den Oxydations- und Desoxydations-Prozeß der Erde,” Zeitschrift für speculative Physik 1, no. 1 (1800): pp. 143–168. See also Schelling, “Vorbericht des Herausgebers,” Zeitschrift für speculative Physik 1, no. 1 (1800): pp. 139–142.

166

E. P. HAMM

combustibles; such as diamond, coal, sulfur, pure metals, and pyrites, and burnt; including earths, salts, and oxidized metals. Combustible minerals are reduced, or de-oxidized, for they attract or draw in oxygen in combustion; those that are burnt are oxidized, or saturated with oxygen. This much is straightforward and follows Lavoisier’s chemistry and is also a restatement of Schelling’s ideas on combustion.36 Steffens then elaborates and gives these basic principles a naturphilosophisch twist by stating that the entire earth is in a continuous state of combustion (curiously, he continues to refer to this as a “phlogistic process” despite the central place he gives to oxygen37), as evidenced by burnt volcanic ejecta that cover large areas, bury entire cities, create islands, and build gigantic craters. The immediate source of volcanic combustion is not to be found in some sort of internal fire (an hypothesis Steffens considered absurd for it required combustion without air to sustain it), but rather in the crater itself, which draws in atmospheric air to sustain itself. Thus great eruptions are accompanied by strong winds in the directions of the volcano. In addition to violent volcanic combustion, the earth is undergoing a continual process of oxidation by the action of air and water. Were these processes to go on unchecked, the earth would eventually be transformed into a burnt lump. Given that “nothing exists in nature without its opposite,” the continuous combustion is accompanied by a continuous reduction, or de-oxidation, which is to be found primarily in coal.38 Chemical and geological evidence shows that coal originates from peat and rich black earth, both of which are, in turn, of organic origin. With this, Steffens has a cycle that makes the inorganic and the organic realms essential to one another. Just as the organic world presupposes an inorganic realm, so also the inorganic world is sustained by a continuous reduction process of the living world.39 Steffens’s geology now was much broader in scope than anything he had laid out in his dissertation, broad enough to encompass all three kingdoms of nature! More than that, Steffens expects to be able to tie the oxidation and reduction cycle to magnetism. He says he will show in another place that “strict laws of the distribution of volcanicity” can “only be determined” through study of the earth’s magnetism and the declination and dip of the magnetic needle, and that the oxidation process and “all of geology (Geologie)…must begin with magnetism”; for good measure he suggests that earthquakes might be an electrical phenomenon.40 Schelling was tantalized by the findings of his Danish disciple, which went far beyond the plans that were laid out in On Mineralogy and the Study of Mineralogy, and he was convinced that they offered the prospect of a “scientific geology” based on Schellingian principles.41 Unfortunately, at least so it seemed at first, the 36

37 38 39 40 41

Schelling, Ideas for a Philosophy of Nature (cited n. 30), pp. 59–64. On the reception of Lavoisier in Germany see Karl Hufbauer, The Formation of the German Chemical Community (Berkeley, CA: University of California Press, 1982). Steffens, “Über den Oxidations- und Desoxydations Proceß” (cited n. 35), p. 149. Ibid. pp. 159–160. Ibid. pp. 159–167. Ibid. pp. 153, 155. Schelling, “Vorbericht des Herausgebers,” Zeitschrift für Spekulative Physik 1, no. 1 (1800): pp. 139–142, at p. 139.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

167

very close collaboration that could only come from being colleagues in the same city could not last. Steffens eventually had to make good on the requirements of his stipend and make his way to Freiberg. He did so reluctantly, despite knowing that the reputation of its Mining Academy was nothing short of sterling, drawing students from across Europe and the Americas, and that Werner’s status in mineralogy was unassailable.42 As it was, the move was very propitious, for his first - hand encounter with Wernerian geognosy allowed him to take up the project for a scientific geology in a way that would scarcely have been likely had he stayed in Weimar. Werner seems the very antithesis of a romantic scientist and there is practically nothing in his published works that indicates any sympathy for Naturphilosophie. In his time Werner was known primarily for his system of identifying minerals and rocks, his classification of minerals and, above all, his classification of rocks. His attention to the proper identification and classification lends itself to characterising him as almost an archetype of the 18th-century taxonomist more interested in collecting for a cabinet than in studying nature in the field, but to do so would be wholly to misunderstand the significance of his projects as they were understood in their time. Werner’s identification system was an extremely fine-grained method for describing mineralogical objects, that is, not just minerals such as quartz, but any kind of rock or inorganic body that might be found. It was difficult to learn as it was based on the five senses, which had to be taught to make a series of many very precise distinctions (there were ten distinct shades of red, for example). But when it was learned, it allowed for accurate standard description of objects, which was of enormous importance for the development of mineralogy and geognosy in the late 18th century.43 Accurate, standardised descriptions allowed for comparisons between observations in various locales without which the development of geology would have been impossible. This system of identification was, and occasionally still is, confused with his system of mineral classification, an error that perhaps has its source in the fact that Werner never published his own mineral classification, though he taught it at Freiberg. In fact, Werner believed minerals should be classified on a chemical basis and his classifications were not unlike those proposed by a number of his contemporaries.44 By contrast, Werner’s classification of rocks was undoubtedly novel. To be clear, minerals are the components which make up rocks as they are found in the world; to be precise, Werner was classifying kinds of Gebirge, or what we might term rock formations. In modern German Gebirge usually denotes a mountain 42

43

44

Steffens, Was ich erlebte (cited n. 11), vol. iv, pp. 202–204; for those who studied at Freiberg see C. G. Gottschalk, “Verzeichniss Derer, welche seit Eröffnung der Bergakademie und bis zum Schluss des ersten Säculum’s auf ihr studirt haben,” in Festschrift zum hundertjährigen Jubiläum der Königlich Sächsische Bergakademie zu Freiberg am 30. Juli 1866, 2 vols. (Dresden: Hofbuchdruckerei von C. C. Meinhold & Söhne, 1866), vol. i, pp. 221–295. Abraham Gottlob Werner, Von den äußerlichen Kennzeichen der Fossilien (Leipzig: S. L. Crusius, 1774). The word fossil was used here in its 18th-century sense, i.e. anything dug out of the ground, and was not restricted to petrified organic remains. There are several published versions of Werner mineral classification. The first authorized publication was C. A. S. Hoffmann, “Mineralsystem des Herrn Inspektor Werners mit dessen Erlaubnis herausgegeben,” Bergmännisches Journal 1 (1789): pp. 369–398.

168

E. P. HAMM

range; Werner’s usage, not unique to him by any means, was in part derived from the terminology of miners, who spoke of being in the Gebirge as soon as they were underground. A Gebirg denoted a rock formation of some massiveness with same position or bed, constructed of the same matter and of a similar origin. Werner’s taxonomy at first consisted of four classes: primitive, which included granite, gneiss, and porphyry; flötz, which included limestone, sandstone, coal, and chalk; volcanic, and alluvial. Later he added a fifth class, the transitional, between primitive and flötz. Historians of geology differ as to the nature of this classification. Some argue, in keeping with Werner’s career at the Mining Academy, that this was a classification concerned primarily with position (or the principle of superposition) and lithology45; others argue that this is an historical classification guided primarily by the time and mode of formation of a rock.46 It is easy to see how such a debate could arise, for Werner wrote that rock formations can be classified into the four groups according to their “nature and origin” and added that one should expect to see in rock formations evidence of different kinds of origin that transpired in the “immense time of the existence of our Earth.”47 He did not, however, specify, what exactly these differences look like, as that was something that would have to be found. In practice, anyone involved with mining would be concerned primarily with position and could see that in this classification primitive rocks would be underneath the flötz. On the other hand, those interested in the origin of rock formations could see that this was a taxonomy of historical objects that developed at a specific time, and that the primitive were the oldest formations. Werner knew that rocks were enormously old but he had no evidential basis for making specific claims about their age and hence did not. He had some evidence about how rocks were formed and he said a little bit about this in his classification, namely that the primitive formations “carry the signs of generation in the wet way.”48 This was terminology taken directly from 18th-century chemistry, which basically had two kinds of analysis: the wet, in solution, and the dry, in fire. This meant that rocks such as granite and other primitives must have crystallized out of solution; the flötz were to some extent chemical precipitates and to some extent mechanical depositions, i.e formed in the wet way; volcanic rocks were formed by fire. Any theory that could account for this taxonomy would have to be deeply informed by chemistry. These questions were not, however, Werner’s concern, for he was developing a taxonomy of rocks, not a theory of the origin of the earth’s crust. 45

46

47

48

Martin J. S. Rudwick, The Meaning of Fossils: Episodes in the History of Palaeontology, 2nd ed., (Chicago, IL: University of Chicago Press, 1985), p. 126. Rachel Laudan, From Mineralogy to Geology: The Foundations of a Science (Chicago, IL: University of Chicago Press, 1987), p. 94. Abraham Gottlob Werner, Kurze Klassifikation und Beschreibung der verschiedenen Gebirgsarten (Dresden: Waltherischen Hofbuchhandlung, 1787), p. 5. First published as an article in 1786, the Kurze Klassifikation is available in an excellent annotated translation, Short Classification and Description of the Various Rocks, edited and translated by Alexander M. Ospovat (New York: Hafner, 1971). Werner, Kurze Klassifikation (cited n. 47), p. 5. For the evidence in support of a “wet” origin of granite see Sally Newcomb, “The Laboratory Evidence of Silica Solution Supporting Wernerian Theory,” Ambix 33 (1986): pp. 88–93.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

169

Unfortunately generations of geologists, including Charles Lyell, turned him into a rash theorist given to the most sweeping generalizations who was too timid to publish his ideas and was surrounded by a protective belt of admiring students who dared not to challenge his authority.49 The debate between Neptunists and Vulcanists was initially and in its most important phase a debate about where basalt belonged in Werner’s taxonomy.50 It was not a battle of worldviews and it is a most blatant distortion that unfortunately still continues to suggest that this debate was somehow guided by fear of offending religious authorities, who of course favoured Neptunism (it is claimed) because it could be reconciled with the Noachian flood.51 The claim that the foundations of Wernerian theory were so ludicrous that it was hard to imagine any responsible geologist accepting them52 misses his taxonomic purpose, nor does it help us to understand why so many students who would become outstanding geologists came to study under Werner, including such towering figures in 19th-century geology as Alexander von Humboldt and Leopold von Buch. Among the students attracted to Werner and Freiberg were a number of leading figures in romanticism, the most outstanding of which was Novalis (Friedrich von Hardenberg). The subterranean world of the mine held many of Steffens’s generation spellbound, but it needs also to be said that Werner’s learning, which was of uncommon breadth even by 18th-century standards, must have played its part as well.53 It should also be noted that in 1800 the Neptunist account of basalt had clearly won the day. Eventually this would change, largely thanks to the work of students of Werner such as Humboldt and Buch, who employed the methods and tools of their teacher to arrive at a different conclusion. In several respects Werner’s work was ideal for Steffens: empirically well founded and with plenty of broader implications, it made no pretense at presenting a theory. Not long after arriving in Freiberg, Steffens found that Werner’s geognosy was becoming ever more important in his work. When he eventually explained his plan to incorporate Werner’s teaching on rock formations to the master himself, the reply was not the one he had hoped for, but pretty much what he expected.54 Undeterred he carried on, with the goal of employing the findings 49

Charles Lyell, The Principles of Geology, [1830–1833] 3 vols., intro. Martin J. S. Rudwick (Chicago, IL: University of Chicago Press, 1990), vol. i, pp. 56–57. See also Alexander M. Ospovat, “The Distortion of Werner in Lyell’s Principles of Geology,” British Journal for the History of Science 9 (1976): pp. 190–198, and Mott T. Greene, Geology in the Nineteenth Century: Changing Views of a Changing World (Ithaca, NY: Cornell, 1982), pp. 19–45. 50 For a helpful discussion of the debate see Otfried Wagenbreth, “Abraham Gottlob Werner and der Höhepunkt des Neptunistenstreites um 1790,” Freiberger Forschungshefte, Reihe C, 223 (1955): pp. 183–241. 51 Werner respected Christianity but his personal position was deism. There is no evidence to suggest that his scientific views were shaped by religious considerations. See Martin Guntau, Abraham Gottlob Werner (Leipzig: BSB B. G. Teubner, 1984), pp. 94–103. 52 Charles C. Gillispie, Genesis and Geology (New York: Harper, 1959), p. 45. 53 Werner developed a classification for an encyclopaedia, Abraham Gottlob Werner Nachlass, Bd. 76, Blatt, 5, Wissenschaftlicher Altbestand der Bibliothek der Technischen Universität Bergakademie Freiberg. See also Henderson, “Practical Knowledge in Romantic Ordering of Nature” (cited n. 9). 54 Steffens, Was ich erlebte (cited n. 11), vol. iv, pp. 227–228. For a general discussion of Freiberg and Werner see ibid., vol. iv, pp. 203–233.

170

E. P. HAMM

of Werner’s geognosy to explain how the earth had formed itself (wie sich die Erde gebildet hat), the task of his Contributions to the Internal Natural History of the Earth, the work that contained “the fundamental theme of his entire life.”55 What about the others who had tried to develop a “chemical geology” for Werner? Steffens dismissed their work in terms that outdo any criticisms made by 20th-century historians of geology. A geology that called for an “original mush” (Urbrey) that was at once simple and yet consisting of distinct parts, that relied on “unexplained laws of affinity,” was built on “the most monstrous presuppositions” and should “forever be banished from science.”56 Despite the judgement in his éloge, Ørsted was clearly taken by the “great discoveries” of the Contributions and its “evolutionary laws of the earth” and incorporated them in several of his works.57 Contributions has two parts, the first of which, “Proof that Nitrogen and Carbon are Representative of Magnetism in Chemistry,” dominates the book. Steffens begins modestly by reviewing a series of recent works in chemistry and concluding that these empirical studies demonstrate that there are two series of earths, siliceous and calcareous, a conclusion, incidentally, that Ørsted had reached independently.58 Steffens then takes these conclusions drawn from the laboratory and aligns them with the very large-scale observations of Werner’s “geognosy.” In his Short Classification, Werner observed in passing that flötz rocks were more calcareous and that primitive rocks were more siliceous. He had already noted then that limestone appeared both as a primitive formation and in the flötz. By 1800, Werner had in his lectures considerably refined his original classification, thanks to observations carried out largely by his students. Though he still held that primitive and flötz formations predominated (volcanic and alluvial deposits were recent and insignificant in comparison), he now had developed the notion of “formations suites” that ran through the primitive and the flötz formations. The slate formation suite included mica-slate and clay-slate in the primitive, and first, second, and third sandstones and sandstone in the coal formation in the flötz. The limestone series began with the primitive limestone and through the first and second limestones, chalk, and marl in the flötz.59 Thus far Steffens’s work is based on 55

56 57

58

59

H. Steffens, Beyträge zur innern Naturgeschichte der Erde (Freiberg: im Verlag der Crazischen Buchhandlung, 1801), p. 93; idem, Was ich erlebte (cited n. 11), vol. iv, p. 286. Steffens, Beyträge (cited n. 55), pp. 80–81. Hans Christian Ørsted, “On the Harmony Between Electrical Figures and Organic Form [1805],” in Selected Scientific Works (cited n. 7), pp. 185–191, at p. 189–188; idem, “First Introduction to General Physics: A Prospectus of Lectures on this Science [1811],” in ibid. pp, 282–309, at p. 305. See also idem, “The Series of Acids and Bases [1806],” in ibid. 227–242, at p. 242. Ørsted’s praise for Schelling was more qualified: the philosopher had “beautiful and grand ideas…but…insufficiently rigorous method,” “Fundamentals of the Metaphysics of Nature Partly According to a New Plan,” [1799] in ibid. pp. 46–78, at p. 77. See also Wilson, “Introduction,” to ibid. pp. xxiv–xxv. Ørsted developed an almost identical series of earths independently of Steffens and, as he carefully pointed out, he did so first. In his éloge for his colleague, Ørsted could find nothing in Contributions that had stood the test of time, but in the first decade of the 19th century priority did matter, see Ørsted, “The Series of Acids and Bases,” in ibid. p. 242. Werner, Kurze Klassifikation (cited n. 47), p. 16; for a discussion of formation suites see the work of the Edinburgh Wernerian, Robert Jameson, System of Mineralogy, 3 vols. (Edinburgh: Constable, 1804–1808), vol. iii, pp. 96–97, and the discussion in Laudan, From Mineralogy to Geology (cited n. 46), pp. 97, 143, 181.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

171

empirical chemistry, particularly the analysis of earths, and empirical geognosy, and it is straightforward and sober-minded by just about any standards. Given his disparagement of Werner in On Mineralogy and the Study of Mineralogy and his interest in Werner only after collaborating with Schelling, one can reasonably conclude that Naturphilosophie did not drive Steffens from empirical science, but rather drove him to a more serious engagement with it. Taking his cue from Schelling, Steffens then developed the siliceous and calcareous series into one set of polarities that are part of a dynamic process governing practically all of nature. The siliceous rocks are linked to a vegetative power (it is in the slate formation suite that plant fossils first appear) that is linked with carbon; the calcareous rocks are linked to animation and nitrogen and, appropriately, it is in the calcareous limestone floetz formations that fossilized fish first appear. Metals are also divided into a fluid, vegetative series (arsenic to mercury) linked with carbon, and a coherent, animated series (antinomy to gold) linked to nitrogen. The fluid series of metals is related to repulsion and the coherent series is related to attraction. Contributions then becomes global, Schellingian physics, arguing that coherent metals gather at the poles, fluid ones in the equatorial regions.60 The second part of Contributions, “In All Organization Nature Searches Only for the Most Individual Formation,” is based largely on the work of Karl Friedrich Kielmeyer (1765–1844) and is little more than an afterthought, a “plain narration”.61 Simpler organisms are governed by irritability, the more complex by sensibility, with everything culminating in man. Contributions has proceeded apace, moving from the simple earths to man, starting with empirical chemistry that is shown to correspond with necessary naturphilosophisch principles, Steffens constructs the world and its development so that it is governed by polarity represented by magnetism, thus fulfilling what he promised to do in “On Oxidation.” Contributions has a slightly misleading title, for it is not in any sense a “natural history” as that term was commonly used around 1800, that is, a description of some aspect of the three kingdoms of nature. It is too speculative for geognosy and it is unlike most other theories of the earth at that time. It is surely geology insofar as it puts forth a unified, theoretical account of the earth’s structure, though it does not look like most other works in the history of geology. Ørsted, for example, gave his enthusiastic approval of Contributions in an article that seems related to Steffens’s project, “The Series of Acids and Bases (1806),” and in two unexpected places: an 1805 article on the harmony between electrical figures and organic form, and in his lectures on general physics of 1811.62 So what kind of work is Contributions? Steffens remarked that everything “corporeal” on our earth represents a “certain stage of development” (Bildung) and that following the progression of this development “constituted the true history of the Earth;” natural history per se is concerned with the external appearance of things, whereas his was an “internal” natural history because it was concerned with the “qualities” that distinguished stages of development from one another. In short, Contributions 60 61 62

Steffens, Beyträge (cited n. 55), pp. 1–17, 29, 56–58, 69, 174–176, 196. Ibid. p. 277. See Ørsted’s works cited in n. 57.

172

E. P. HAMM

bears only some resemblance to older “theories of the earth.”63 It falls between the cracks of natural history, theories of the earth, and speculative philosophy; it fits neither our categories nor those that prevailed in 1800.64 Is Contributions a history of the earth? Yes and no is the only answer to such a bald question. Contributions does reconstruct the earth’s development over time, of that there can be no doubt. Contrary to those who would argue (and have) that Naturphilosophie was concerned with the logical relations between things, in this case formations or strata, and that the Wernerians had a “willful neglect” of chronology, Steffens reported that Werner himself was seeking to perfect his system by studying the relationship of species of fossilized animals “to their geognostic age.”65 To accuse someone working in 1800 of neglecting chronology is to expect something the evidence could scarcely sustain. Except for recent alluvial and volcanic depositions that were recorded in written testimony, the tools for the absolute chronology of sedimentary formations were only beginning to be developed. Here some distinctions are necessary, for the kind of history that Steffens was doing is best understood as a genetic history, a history of development unfolding according to carefully set out laws. This was something akin to what James Hutton did in his Theory of the Earth, a work that despite positing a very ancient earth is not historical, for it posits laws at work in an endless cycle of events. Steffens’s theory, on the other hand, was directional. This does not mean that every single geological event was somehow preordained, but rather that the pattern of development has a certain direction that becomes increasingly clear over time. All of development is aimed at human development. The best statement of what Steffens was trying to do in Contributions can be found in a remarkable passage from Ørsted’s “Reflections on the History of Chemistry” of 1807. Although there is no direct reference to Steffens, the passage is an excellent summary of the story told in Contributions: By studying the history of one’s own science in this way one acquires insight into the development of the entire human spirit. Not just chemistry but all human knowledge has always, although with varying clarity, intervened in the essence of things. This has always been developed through a continually renewed struggle, which, however, has resolved itself in perfect harmony. And it is not just science, not just human nature, it is all of nature that develops according to these laws. To show this to its full extent would be to present a complete natural science and a complete history. Therefore…I am forced to content myself with a simple account of a single observation. The evolution of the earth seems to me to be the most suitable for this. We are able to penetrate into the darkness which envelops the history of our globe as we penetrate into its bowels and compare the deeper layers with the older and the newer ones. By examining these layers and the fossilized or imprinted creatures which are found in them, we learn that the earth began with enormous creative forces but with little definite direction. Through alternating expansions and contractions, it has gradually killed and buried its earlier creatures in order to make room for the present chain of creations with man at the head. It is clear to every open-eyed observer of nature that the creative and organizing forces have alternated, with an ever increasing

63 64

65

Steffens, Beyträge (cited n. 55), p. 96, italics in original. See Steffens, Was ich erlebte (cited n. 11), vol. iv, pp. 292–294 for some of the varied reactions to Beyträge. Steffens, Beyträge (cited n. 55), p. 86, and on p. 87 where Steffens writes approvingly of Georges Cuvier’s palaeontological work in the Paris basin.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

173

predominance of the organizing ones, however, and that it has arrived at the stage of development at which it now stands only after many struggles. In short, the development of the earth is like that of the human spirit. The harmony between nature and spirit is hardly accidental.66

Human history and the history of nature are bound together unfolding as they must, and that is the key. Unlike the history that geologists would develop in the first decades of the 19th century, this was not the reconstruction of contingent events, it was the working out of necessary laws. In a letter to Goethe, Steffens observed “that arbitrariness in science is detested above all else.”67 Arbitrary or contingent events are the very stuff of geological and human history; they have no place in the Naturphilosophie of Schelling, Steffens, and Ørsted, all of whom saw the earth as the product of dynamic chemistry, not contingent history. Steffens, it has to be said, soon extended his laws of the earth’s development far beyond the boundaries of reason. In his Foundations of Philosophical Natural Science he extended the polarities to encompass a doctrine of race. The black race was the southern principle, the Mongolian the northern, the Malaysian was the eastern, and the American the western. The Blacks and Mongolians have no trace of history; the Malaysian and American have only a vegetative history, that is they are prehistoric and mythological. The Caucasian race is temperate, historical, and has the stages of development in itself.68 Goethe had gone to hear Steffens lecture on his Foundations at Halle (where he was teaching in 1806) and, from behind a curtain where he sat so as not to be distracting, he did not think it was so bad. Steffens, like Werner, was a famous teacher, though more of an orator than Werner. After Goethe had read Foundations he wrote Friedrich August Wolf, the distinguished classicist at Halle, that the most polite thing to be said of the book was that it had a “Preface that was sweet as honey, but we lay people must gag on its contents.”69 What then became of the chemical construction of the earth developed by Steffens, using the resources supplied by Schelling and Werner, and endorsed by Ørsted? Historians of science have tended to concentrate their attention on historical geology and palaeontology, sciences that reconstruct the earth’s history on the basis of local evidence painstakingly gathered. However, the importance of chemistry in mineralogical and geological explanations lasted long after Werner.70 Consider the example of Alexander von Humboldt, in particular his treatment of 66

67

68

69

70

Ørsted, “Reflections on the History of Chemistry, a Lecture,” in Selected Scientific Works (cited n. 7), pp. 243–260, at p. 259. H. Steffens to Johann Wolfgang Goethe, 3 September, 1806, Briefe an Goethe, ed. Karl Robert Mandelkow, 2 vols. (München: Deutscher Taschenbuch Verlag, 1988), vol. i, pp. 454–455, at p. 455. H. Steffens, Grundzüge der philosophischen Naturwissenschaft. Zum Behuf seiner Vorlesungen (Berlin: Realschulbuchhandlung, 1806), pp. 196–197. Goethe to F. A. Wolf, 31 August, 1806, Goethes Werke, herausgegeben im Auftrage der Großherzogin Sophie von Sachsen, 4 parts, 133 vols. in 144 [1887–1919] (München: Deutscher Taschenbuch Verlag, 1987), pt IV, vol. xix, pp. 186–188, at p. 187. On chemistry in geology see the helpful discussion in Laudan, From Mineralogy to Geology (cited n. 46), chs. 3, 5, 7, and 8, and Bernhard Fritscher, Vulkanismusstreit und Geochemie: Die Bedeutung der Chemie und des Experiments in der Vulkanismus-Neptunismus-Kontroverse, Boethius; 25 (Stuttgart: Franz Steiner, 1991).

174

E. P. HAMM

volcanoes. Humboldt asked, what burns in volcanoes? The answer was not unlike that Steffens gave in “On Oxidation”: it was the metal oxides. More than that, Humboldt was looking to establish laws on which one could base the progressive development of the earth. Volcanoes, in particular their distribution throughout the world, was not accidental. They were linked by some sort of internal chemical process.71 More striking is the case of Leopold von Buch (1774–1853), one of the most distinguished geologists in the first half of the 19th century. Buch was among the first of Werner’s many students to openly abandon Werner’s explanation of the origin of basalt. One of the things that made Buch famous was his analysis of dolomite. Déodat de Dolomieu (1750–1801) had shown in 1791 that the white mountain peaks around the Fassa valley in the southeastern Alps were not marble. Simple chemical tests showed the rock did not effervesce as marble and other calcareous rocks do when a weak acid is poured on them. The predominant earth in this rock, which became know as dolomite, was magnesia. Buch took this simple chemical fact and aligned it with another: magnesia is an important part of augite porphyry. He argued that what was now dolomite had originally been deposited as calcareous limestone, which was “dolomitized” by the intrusion of augite porphyry and magnesia vapours. Buch then claimed that augite intrusion and dolomitization was responsible for creating the Alps. Chemistry works everywhere, so Buch generalized his theory to explain the formation of all mountains.72 Buch was anything but a Naturphilosoph. Steffens, he wrote in 1804, made him want to cry out “O BACON!… What has become of the way of experience? I fear the truth will have vengeance on him [Steffens]. So much intellect and acumen wasted!”73 His case is telling for it shows that the chemical tradition was leaving its mark on geology well into the 19th century. It is also telling for another reason. In 1806, Buch gave his inaugural address to the Berlin Royal Academy of Sciences, “On the Progress of Formations in Nature.” In this speech Buch is easily the equal of Steffens when it comes to speculation. Buch constructs the world not from chemistry, but from its offshoots, mineralogy and crystallography. Every formation in the world, from a simple crystal or moss to mountain chain or human form, is the product of a “conflict of forces,”74 which gives rise to form in previously undifferentiated matter. The most fundamental of these forces is that which 71

72

73

74

Alexander von Humboldt, “Über den Bau und die Wirkungsart der Vulcane in verschiedenen Erdstrichen,” Abhandlungen der Königliche Akademie der Wissenschaften zu Berlin aus dem Jahren 1822– 23 (Berlin: Königliche Akademie der Wissenschaften, 1825), pp. 137–156, pp. 144, 153. Laudan, From Mineralogy to Geology (cited n. 46), pp, 194–197. Buch’s theory did not last, in part because of chemistry. Jakob Berzelius argued there were no such things as magnesia vapours for the simple reason that magnesia could not sublimate. Leopold von Buch letter to Karsten, 17 May 1804, in Buch, Briefe an D .L. G. Karsten. Zu seinem 150. Geburtstag, ed. Julius Schuster and Robert Bloch (Berlin: W. Junk, 1924), p. 26, emphasis in original. Later Buch was slightly more generous: “As always there is lots of truth in it [Steffens’s work] mixed up with half-baked ideas.” Letter to Karsten, 27 January 1805, ibid. p. 29. Leopold von Buch, “Ueber das Forstschreiten der Bildungen in der Natur,” in Gesammelte Schriften, edited by J. Ewald, J. Roth, H. Eck, and W. Dames, 4 vols. (Berlin, 1867–1885), vol. ii, pp. 4–12, at p. 6 and passim. Steffens was aware of how closely Buch’s ideas followed his own, Steffens, “Ueber das Verhältnis der Naturphilosophie zur Physik unserer Tage,” in Alt und Neu, 2 vols. (Breslau: J. Max, 1821), vol. i, pp. 67–84, p. 77.

STEFFENS, ØRSTED, AND THE CHEMICAL CONSTRUCTION OF THE EARTH

175

creates crystals, and the most fundamental conflict is between it and the force of gravity. Sounding remarkably like Goethe, Buch announces that everything begins with the formation of granite.75 He then works his way through the Wernerian formations, duly noting the appearance of fossils (also the product of a conflict of forces), particularly the massive numbers of them in calcareous limestone. Following the lead of Georges Cuvier (1769–1832), he notes the frequent catastrophes that have wiped out life; following the lead of Lamarck he notes the new forms of life that come into being. The physical formation (Bildung) of humans, a recent event in the history of the earth, is also the product of the conflict of physical forces. Everything culminates in the cultural Bildung of humanity. Buch concedes that he is offering a promissory note rather than anything scientific, but he does say he sees “geology” as the science that might succeed in finding the laws governing the development of everything. These accounts of Bildung, chemical, geological, and crystalline, are in some important ways frozen out of history. Steffens and Buch, at least for a moment (just as Napoleon was bearing down on Prussia), were both deeply aware that the Bildung that concerned them was a temporal process, but it was a process governed by inherent, internal polarities or forces that were working themselves out and unfolding in the world. It should not be wondered that the generation of German Romanticism, be it Naturphilosophisch or otherwise, was absorbed, and selfabsorbed, with the notion of Bildung, for it had been inscribed into the structure of the earth. It was also inscribed in the self: “Do you want to know nature? Cast a glance inside yourself and in the stages of intellectual development (Bildung) you may be granted the privilege of seeing the stages of nature’s development. Do you want to know yourself ? Study nature and her deeds are those of your mind.”76 York University

75

76

Goethe’s essays on granite were unpublished in his lifetime, see “Granit I [1784]” and “Granit II [1785],” in Johann Wolfgang Goethe, Sämtliche Werke: Briefe, Tagebücher und Gespräche, 40 vols. (Frankfurt a. M.: Deutsche Klassiker Verlag, 1985–1999), vol. xxv, pp. 311–316. The irony here is that the theory of mountain formation that Buch would develop was absolutely anathema to Goethe. It was the sort of violent change that hindered peaceful Goethean Bildung. Steffens, “Ueber die Vegetation,” in Alt und Neu (cited n. 74), vol. ii, pp. 36–109, at p. 102.

THE CULTURE OF SCIENCE AND EXPERIMENTS IN JENA AROUND 1800 OLAF BREIDBACH

1. INTRODUCTION 1.1. Local traditions Around 1800, Jena University was at the centre of one of the highlights in European science. Jena was a small town at that time, and a pedestrian could cross it in less than a quarter of an hour. It was, in fact, a poor area to live and work in. (Steiger 1958; Walther, H. G. 2001). As has been shown by recent research, the economic conditions during the interregnum of Anna Pawlowna were anything but ideal. Intellectual life was also limited (Berger 2001). Even the magnificent friend of Johann Wolfgang Goethe, the duke Carl August von Sachsen Weimar Eisenach, was mostly engaged in hunting and military projects, and was no philosopher at all. Jena was neither Paris nor London, Vienna nor Goettingen. Nevertheless, it became prominent in literature. The university, in spite of its small size, was thriving, and it might even be that it was a profitable one for the district Sachsen-WeimarEisenach. However, students were attracted not only by the philosophers Hegel and Schelling, but also by the physicians Loder and Hufeland. As Rasche has shown, around 1800, more than one-third of all students in Jena (from more than 800) matriculated in the faculty of medicine (Rasche 2001). By number of its students, Jena was at the fourth position within the German universities in Protestant areas. In terms of physical size and resources, however, Jena University was one of the poorer universities in Germany around 1800 (Hammerstein 2001). In contrast, the not-too-distant university and academy of sciences in Goettingen were well established: the libraries were magnificent, and the collections gathered by people such as Lichtenbergor—in the history of nature—Blumenbach, were significant (Walther, G. 2001). Such outstanding structures were lacking in Jena (Zaunstöck 2001). The Jenensian Naturforschende Gesellschaft’ did not even possess a philosophical laboratory but was situated in the private house of the botanist August Johann Georg Karl Batsch (1761–1802) (Ziche 2001). The library, which formed the nucleus of a library of science in Jena, possessed about 796 volumes (Ziche et al. 2000). The natural history collection, supervised by the mineralogist Lenz, was unsystematic (a Kunstkammer) 177 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 177–216. © 2007 Springer.

178

O. BREIDBACH

(Krehr-Hartmann 1999). However, the botanical garden, founded 1794, was organized according to the new natural system developed by Batsch, under direct influence of Goethe (Jahn 1963; Poliansk. 2001). Despite its small institutions, the University of Jena, advertising in the Jenensian Allgemeine Literaturzeitung (1790 and 1805) claimed that with such institutions it could provide exceptional support for students in their studies of the natural sciences.1 It has been argued that this town, with its great Kantian tradition and the backing of the genius Goethe, was nothing less than the birthplace of the German scientific culture, which the Prussians claimed changed the structure of science in the 19th century. This view even found its icon in a sketch of the meeting between Goethe and Schiller, the “happy event.” The two were later joined by Alexander von Humboldt and his brother Wilhelm. That often cited picture, reproduced as a print by L. J. Aarland adopted from A. Müller, is a fantasy, but it describes a certain attitude towards Jena around 1800. According to that view, Jena provided a context for the development of German romanticism, a product of the close interaction of a particular group of scientists, philosophers, and poets. What is behind that fantasy? Alexander von Humboldt stayed in Jena, he met the romantic scientist Ritter, who had neither a doctorate nor any internationally recognised publication to his name at that time, and asked him to contribute to his (Humboldt’s) volume on animal electricity (Humboldt 1797). Fichte had taught there, delivering his famous lectures to the German nation. In 1799 he was forced by the government to resign. Yet, already by around 1800, his disciple Schelling, who taught in Jena, had earned at least a national reputation. People like Herder and Wieland worked just 20 km away, in the small residence town of Weimar. A list of names of prominent persons working in the humanities and sciences in the area of Weimar/Jena would include many outstanding thinkers of the time—even genii in statu nascendi. Did these different authors reflect a common tendency in their studies, thinking, and publications? Or were the Jena/Weimar poets and philosopher-scientists an example of a specific local culture? An analysis of their styles of scientific thinking allows us to answer this question. By describing the specific characteristics of the Jenensian tradition we can reflect on the diversification of scientific practice and tradition, and the development of scientific thinking around 1800. The picture of the lucky event cheering just a few key figures can be identified as a retrospective construction. There were many interactions, and a complex of various research traditions, that contributed to the flowering of Jena University and its surroundings around 1800. The culture of science in Jena around 1800 cannot be reduced simply to romanticism. The following list, (Fig. 1) adapted from a handout of the Sonderforschungsbereich 482—Ereignis Weimar Jena—Kultur um 1800, illustrates the number and range of persons active at the time.

1

Intelligenzblatt der Allgemeinen Literatur-Zeitung (i.F. ALZ) 1790: No. 1, 1805:. No. 37, 39.

THE CULTURE OF SCIENCE AND EXPERIMENTS Anna Amalia (1739–1807) August Johann Georg Carl Batsch (1761–1802) Friedrich Justin Bertuch (1747–1822) Johann Joachim Christoph Bode (1730–1793) Karl August Böttiger (1760–1835) Clemens Brentano (1788–1842) Sophie Brentano (1770–1806) Carl August (1757–1828) Carl Friedrich (1783–1853) Johann Wolfgang Doebereiner (1780–1849) Johann Peter Eckermann (1792–1854) Friedrich Hildebrand von Einsiedel (1750–1828) Johannes Daniel Falk (1768–1826) Carl Ludwig Fernow (1763–1808) Johann Gottlieb Fichte (1762–1814) Jakob Friedrich von Fritsch (1731–1814) Friedrich Wilhelm August Fröbel (1782–1852) Carl Friedrich Ernst Frommann (1765–1835) Ludwig Friedrich von Froriep (1779–1847) Luise von Göchhausen (1752–1807) Johann Wolfgang von Goethe (1749–1832) Karl Gräbner (1786–1845) Johann Diederich Gries (1775–1842) Johann Jakob Griesbach (1745–1812) Georg Wilhelm Friedrich Hegel (1770–1831) Johann Gottfried Herder (1744–1803) Johann Gottlob Heynig (1772–1837) Friedrich Hölderlin (1770–1843) Christoph Wilhelm Hufeland (1762–1836) Alexander von Humboldt (1769–1859) Wilhelm von Humboldt (1767–1835) August Wilhelm Iffland (1759–1814) Christian Joseph Jagemann (1735–1804) Jean Paul (1763–1825) Johann Georg Keil (1781–1857) Martin Gottlieb Klauer (1742–1801) Ernst August Friedrich Klingemann (1777–1831) Carl Ludwig von Knebel (1744–1834) August von Kotzebue (1761–1819) Georg Melchior Kraus (1737–1806) Jakob Michael Reinhold Lenz (1751–1792) Johann Georg Lenz (1745–1832) Justus Christian Loder (1753–1832) Maria Pawlowna (1786–1859) Sophie Mereau (1770–1806) Johann Heinrich Meyer (1760–1832)

179

180

O. BREIDBACH

Johann Carl August Musäus (1735–1787) Novalis: Friedrich von Hardenberg (1772–1801) Lorenz Oken (1779–1851) Heinrich Eberhard Gottlob Paulus (1761–1851) Carl Leonhard Reinhold (1758–1823) Friedrich Wilhelm Riemer (1774–1845) Johann Wilhelm Ritter (1776–1810) Johann Friedrich Rochlitz (1769–1845) Friedrich Wilhelm Josef Schelling (1775–1854) Friedrich von Schiller (1759–1805) August Wilhelm Schlegel (1767–1845) Caroline Schlegel–Schelling (1763–1809) Dorothea Schlegel (1763–1839) Friedrich Schlegel (1772–1829) Carl Christian Erhard Schmid (1761–1812) Johanna Schopenhauer (1766–1838) Friedrich Gottlob Schulze (1795–1860) Christian Gottfried Schütz (1747–1832) Johann Samuel Gottlob Schwabe (1746–1836) Henrik Steffens (1773–1843) Charlotte von Stein (1742–1827) Ludwig Tieck (1773–1853) Christian Gottlob von Voigt (1743–1819) Johann Heinrich Voß (1751–1826) Johann Heinrich Voß d. J. (1779–1822) Christian August Vulpius (1762–1827) Christoph Martin Wieland (1733–1813) Ludwig Wieland (1777–1819) Johann Karl Wözel (1765–1836) Pius Alexander Wolff (1782–1828) Caroline von Wolzogen (1763–1847) Fig. 1. Selection of persons active, in Jena/Weimar around 1800 – adapted from a handout of the Sonderforschungsbereich 482 – Ereignis Weimar Jena – Kultur um 1800. 1.2. General trends in the situation about 1800 Herder and Wieland, and the Kantian school of philosophy in the University of Jena, can be regarded as the culmination of the German enlightenment tradition (compare Henrich 1991; Hinske, Lange, and Schröpfer 1994; Manger). Science, as had been practised in Goettingen—by Georg Christoph Lichtenberg (1742–1799), for example—also followed this tradition. Already, in the decades before 1800, the style of thinking in Goettingen had shifted in this direction. Kantian philosophy is analytic. Kant praises mathematics as the ideal form of science. Thus, Kant and his explicit followers are far from a speculative approach to the idea of nature as proposed by Schelling in the last decade of the 18th century. Nevertheless, in

THE CULTURE OF SCIENCE AND EXPERIMENTS

181

1778, when giving a lecture in the academy of sciences to present a new method for exploring the nature and motion of electrical matter, Lichtenberg spoke in analogies, describing the earth as a gigantic tourmaline, which gave rise to those polarities which later “electrified” the community of natural philosophers around Schelling. Obviously, the idealistic vocabulary, and the concepts connected with it, which had been characterized as alchemical ones by Ghiselin (2000), are neither specific to the Jenensian locality nor a domain of Schellingian or Hegelian philosophies. These traditions arose independently from the discussion but nevertheless still form part of the Jenensian Naturphilosophie. Any general statement describing the period of romanticism, and trying to connect with this term certain trends in places such as London, Copenhagen, Jena, and Heidelberg, must also deal with specifics if it is to enhance our understanding of the culture of science around 1800. Thus this study does not follow the idea to describe factors that may allow to qualify the extraordinary situation in Jena described as the lucky event but makes use of the extensive knowledge we had from this place to form a case study. For an adequate understanding and evaluation of the common characteristics of scientific discourses, a closer analysis of the local cultures is necessary. This means considering the cultural, social, and economic situation of each region, as these provided the context within which scientific developments took place. Factors such as scientific resources, equipment, and technical and experimental traditions, are also relevant. In Jena, such a research project has been running under the heading “the event Weimar/Jena—culture around 1800,” since 1998. This report tries to summarize some of the results from the project, and what follows is a first sketch of the local culture of science. This paper is therefore an examination of local interactions. 1.3. Aims and scope The present study will describe some preliminary results in study of the culture of sciences in Jena University around 1800, ending with a short sketch of the tradition of natural philosophy in Jena in the first decades of the new century. It will comment on the reactions of the prominent Jenensian antagonists to that movement: the philosopher Jakob Friedrich Fries (1773–1843) and the botanist Matthias Jacob Schleiden (1804–1881). For ongoing discussion, it should be noted that the logician, Gottlob Frege (1848–1925), and the evolutionary biologist, Ernst Haeckel (1834–1919), also resided at Jena University at the end of the 19th century. However, the present study is restricted to the decades from 1790 to the 1820s, and describes the programme of lectures, texts, and some of the publications of that period in order to provide an idea of the style of Jenensian scientific education and scientific practice. Thereby, one has to consider, that part of the education and research in natural sciences was done by persons employed in the medical faculty. Thus, it will be necessary not only to describe the situation in the philosophical faculty where the natural sciences were established but also to look at the medical faculty to determine whether there was a particular Jenensian education practice in that area. Thereby it will be necessary to compare educational programmes and eventually the research done by persons in these different faculties. At present such a

182

O. BREIDBACH

comparison can be only an exemplary one. The paper will compare the work on electricity published by two persons, the professor of medicine, Franz Heinrich Martens (1778–1805), and the natural scientist Johann Carl Fischer (1763–1833), who—teaching in Jena—wrote a dictionary of physics, and the first monograph in the German language about the history of the physical sciences that I know of. This will be paralleled by a short survey of the development of the Naturphilosophie (philosophy of nature) after Schelling, and a consideration of Schelling’s influence on the educational programme of the philosophical faculty after 1799. I will argue that there was in fact a particular style of German Naturphilosophie connected with Jena University. This style, however, was not reflected in the educational programme in the sciences. Finally, this study will address the questions of experiments. The famous ones were the self-experiments done by Ritter and Humboldt, and Ritter’s discovery of what is today called ultraviolet radiation in 1801. We have studied that experimental practice in detail, even rebuilding part of the equipment and redoing some experiments. We have also been engaged with a particular instrument of physics, the electric machine built in Weimar before 1800. These studies have given us indications of experimental practice in central Germany around 1800.

2. CASE STUDY 2.1. Natural history This study will refrain from explicitly commenting on the principal aspects of scientific development following the proclamation of the end of natural sciences by Lepenies in 1976.2 It would be useful, however, to consider some general developments. Lorenz Oken (the former Jenensian scientist and philosopher, cf. Breidbach 2001b), for example, published his magnificent and well-received natural history for the people (Naturgeschichte für alle Stände) in the 1830s. Even Charles Darwin (1809–1882) worked in this tradition of natural history (Richards 1992). Indeed, natural history was not out of vogue until the end of the first half of the 19th century, and remained a prominent branch of the sciences until that time. Central Germany, including Goettingen, Jena, Leipzig, and Weimar, was one of the strongholds of that tradition. Its scientists provided a significant contribution to the main reference books for education in natural history, in Italy, for example, at the University of Padua, although the work of the scientists at the Parisian museum d’historie naturelle predominated (Breidbach and Frigo 2003). What impact did this speculative philosophy of nature (Naturphilosophie) have on the development of science in general? It has been argued that the rise of the Schellingian school of Naturphilosophie pointed to a transitional 2

It would be unfair to reduce Lepenies’ important contribution to that aspect which is presented by the title of his book. He saw a discontinuity in the development of science, replacing a historical descriptive by an analytical inductive approach. Thereby for him, the end of natural history is an example for the formation of new ideas to enface temporal dynamics.

THE CULTURE OF SCIENCE AND EXPERIMENTS

183

phase in the alteration from a science of nature embedded in natural history, to an analytical science, as espoused by the Jenensian botanist Schleiden in the 1840s (Schleiden 1844; Breidbach 1988, 1998a). What is also evident, however, at least in Germany, is the persistence of natural history, and even the extension of its influences on science in general, up to the 1850s. That influence was largely in the fields of functional morphology in medicine and the biosciences which were formed out in a typological tradition. This tradition developed out of the concept of morphology as espoused by Goethe and, as a result of his influence, by scientists associated with Jena University (Kuhn 1988; Wyder 1998; Breidbach 2001a). Lorenz Oken, a follower of Schelling, is famous for his discourse with Goethe about the identification of skull segmentation and his political engagement in the liberal movement in Germany in the early 19th century. In so far he stands for both the continuation of the ideas of a natural history and for the impact of Schelling on that science. What Oken did was a reconsideration of the order paradigm of former natural history. The intensive discussion about a “natural system” in biological systematics was in a critical situation around 1800 (Dieckmann 1992; Stevens 1994; Müller-Wille 2000). Different solutions to the problem had been proposed, but neither could secure the order categories proposed. In this situation the idea of a typology of nature seemed to establish a conceptual framework to deal with that problem. Goethe was pursuing this idea especially in regard to botanical classification and plant morphology (actually the term “morphology” was coined by him). In his Naturphilosophie Schelling provided an argument to deduce a certain order paradigm for a systematization of natural entities (Breidbach 1986). He did not detail this argument, however. For zoology and botany, this was done by Lorenz Oken (Bach 2001c). Hereby Oken employed the essentials of Schelling’s schemes. In the effect Oken transferred philosophical principles into natural history. At least for the first half of the 19th century, he was quite effective in doing so (Breidbach 2001b). A prominent figure in the development of 19th century German science, Oken was the editor of Isis, an influential German science journal, and instrumental in the formation of the German Society for Scientists and Physicians, a social structure later copied by the British Association for the Advancement of Sciences. Through people such as Lorenz Oken, speculative Naturphilosophie was a prominent part of the scientific dialogue of the first half of 19th century. Accordingly, one has to reconsider any idea of periodization employing simplistic categories such as romanticism, history of nature, or the natural sciences (Naturwissenschaften). Here, it will be argued that this Naturphilosophie started as a philosophy of science, around 1800. In the generation of younger philosophers following Schelling it evolved into a new kind of metaphysics, separating itself from the sciences (Breidbach 2000). The philosophical arguments of this new generation of philosophers were removed from the sciences, whereas the original Schellingian influence had become effective via scientists like Oken (Breidbach and Ghiselin 2003). That influence was not restricted to German science. It applied— especially in relation to the work of Richard Owen—also to the biological sciences in England (Rupke 1994).

184

O. BREIDBACH

2.2. Sciences From Stichweh’s (1984, 1994) analysis of the professionalization of science in Germany, we know that a reorganization of physics and chemistry in education was undertaken before 1800. This is reflected in the late 19th century Jenensian situation. The term “natural science”—Naturwissenschaften—was introduced as a classificatory term in the organization of the disciplines in the philosophical faculty of Jena University in 1790 (Ziche 1998). Thus at least some parts of a new terminology for the analytical sciences had been formed before the onset of Naturphilosophie and apart from any influence of scientific romanticism. In the classification that was a consequence of the French reorganization of universities in the beginning of the 19th century, that classification was even more explicit. There natural history and the analytical sciences were taught in two different faculties. The post-revolutionary University system was organised into analytical and applied sciences, combining mathematics, physics, fortification, astronomy, navigation, chemistry, and engineering, while the disciplines of zoology, botany, and mineralogy were located in the medical faculty (Breidbach, Frigo, and Piovan 2003). In the 18th century, natural history was introduced as a descriptive science, aimed at a systematisation of the entities of the realm of nature. Originally, physics and chemistry were also referred to as descriptive sciences. In fact, Lavoisier’s work in chemistry resulted in a new systematics and—accordingly—in a new terminology. The difference of physics and chemistry vs. natural history, was not such an obvious one. In fact, until the last decades of the 18th century these two traditions—the experimental and descriptive sciences—were classified as subparts of a phenomenology of nature, the former natural history. Physics, on the other hand, was seen as a science closely associated with applied mathematics (Knight 1981, 1984). The nucleus of the new experimental physics around 1800 developed from studies of galvanism, magnetism, and electricity (Frercks 2003). These led to important methodological advances of the experimental and inductive sciences. Specific measuring devices and experimental set-ups were devised to study these new types of phenomena that could not be detected by our human sensory equipment. These experimental approaches were then transferred to the study of physiological phenomena. While around 1760, physiologists such as Haller had aimed at classifying the types of animal reactions,3 scientists now looked for an understanding of the principles from which such reactions were invoked. As direct sensory control of the phenomenon was not possible, experiments had to be designed simply to protocol the various types of electrical phenomena. The standard test device was simply an organic tissue, the leg of a frog (Pera 1992). The reaction of this tissue demonstrated the presence or absence of galvanic activity during the experiment. Now the question was, whether the contraction of the frog’s muscle in response to electrostimulation was an artefact, or indicated the presence of galvanic 3

See the introduction in Roe (1981).

THE CULTURE OF SCIENCE AND EXPERIMENTS

185

current actually produced by the animal tissue. In the 1750s, electric fish had been discovered in the Amazon (Steffens 1818). Apart from this phenomenon, their significance was restricted and only a small group of scientists had actually seen the fish. Furthermore, no measuring device was available at that time to actually demonstrate that the fish produced electricity. 2.3. Science and Naturphilosophie The core of all the studies on electricity was a proper definition of the phenomenon. Experiments were designed just to compare the different types of electricity, such as that produced by an electric machine, a Voltaic pile or by organic tissue. The Jenensian scientist Johann Wilhelm Ritter (1776–1810), together with Alexander von Humboldt (1769–1859), tried to experiment with these phenomena in order to qualify the effects of electricity on organic tissue, and to determine the extent to which the body simply experiences such electricity, or produces it (Wiesenfeldt 2003). The problem was that there was no established theory for understanding the processes by which electricity was formed (Moiso 1994). Nor was it easy to see that different modes of production of electric phenomena actually resulted in an identical product. Electricity was known only by its effects on various substances, on animals, and man. Classifying the phenomena proved difficult. Such a classification would have to fit into a general theory of the natural world. Such a general theory was provided by Schelling, who believed that the experiments done by Ritter had revealed something akin to the principal processes of nature (Breidbach 1999b, 2003). Schelling developed a vocabulary which set at those processes, which was subsequently adapted by Ritter. Schelling offered him a scheme to systematise his experiences. In these two scientists, the interaction between the new Naturphilosophie and experimental science can be studied (cf. Lenoir 1978, 1982). Ritter worked in a terminological framework adapted from Schelling. And Schelling structured his approach to Naturphilosophie according to experiments performed by Ritter (Breidbach 2003). For the history of German idealistic philosophy this is of utmost relevance. Naturphilosophie can be demonstrated as having been part of the scientific discourse around 1800 and not as having been set apart from that discourse. In an analysis of the development of science an analysis of that situation also is of relevance. It is to ask how far, the adaptation of Schellingian philosophy actually was of relevance for the design of Ritter’s experimental approach. A case study, focusing on the interactions of Ritter and Schelling, thus, might allow a better understanding of the interaction of experimental practice, theoretical frameworks, and the reception of such an interaction, both at the time around 1800 and during the following decades of scientific development. Ritter was not part of the academic staff of Jena University (Richter 1988). He was guaranteed the privilege of giving one lecture on animal galvanism by the government of Weimar. This lecture, however, was never officially announced by the university authorities, and his ideas seem not to have been essential for the educational programme of the university.

186

O. BREIDBACH 3. ACADEMIC LECTURES

3.1. Overview The Jena University has already been described as one based on assistant professors (Müller 2001a). This was so even in the sciences. This meant that few persons actually held a chair: They were Laurenz Johann Daniel Suckow and his successor Johann Heinrich Voigt in physics, and the botanist Batsch in natural history. In chemistry most lectures were performed by the associate professor (Extraordinarius) Johann Friedrich August Göttling (Fig. 2). In addition there were a series of assistant professors or lecturers who were engaged for a shorter period of time. As can be seen in Figures 2, 4, and 5, these scientists were engaged in university education for a period of only several semesters. They did not deliver special courses but contributed to the basic educational programme. As can be seen from the archive material, even for the lecturers it was a most attractive situation to join, because firstly, it documented the qualifications to teach the essential stuff of their disciplines, and secondly, it […] guaranteed a minimum number of students whose fees guaranteed a degree of financial security as the lecturers did not get any salary from the university. The rich archive material allowed a very detailed reconstruction of the study programme in the sciences. Interestingly, a specific touch of romanticism was not present in that programme. Instead, the standard scientific education program (Stichweh 1984; Lind 1992) was offered. To secure such a somehow “negative” statement we analysed the complete Jenensian programme of lectures in the sciences (Bach and Breidbach 2001). In the years from 1790 to 1807, there were 334 lectures devoted to Naturwissenschaften. In the period between 1790 and 1807 the number of lectures offered were highest in chemistry, 2nd botany and 3rd physics. In chemistry, the number of lectures delivered grew between 1790 and 1796 from 5–9 per year, thereafter it was declining. Only the lectures in the general natural history showed a constant decline from 1793 to 1807. In 1807, however, there were 4 lectures a year, which matched the situation of 1798. Within physics and in the areas of zoology, botany, and mineralogy, the lectures announced followed a constant scheme. In chemistry, there was always a general course available. Between 1790 and 1796 lectures on popular chemistry were announced. In addition, special courses on the history of chemistry, systematic chemistry, chemical technology, and about certain chemical substances such as gases or salts were announced. Some of these were delivered only once, whereas the general course on chemistry was given in each semester. Likewise in botany and mineralogy there were constant offers in general courses with additional special lectures on various systematic matters, or, within botany, the offer of special study trips. There was just one lecture on the actual subject of plant sexuality, but there was always a lecture devoted to the general history of nature. In physics, standardisation was even more obvious. There was a lecture on theoretical and experimental physics which was replaced from 1799 to 1803 by a lecture on experimental physics. (The latter course already had been given in

THE CULTURE OF SCIENCE AND EXPERIMENTS

187

Appendix 1 Persons Batsch, August Johann Georg Carl** Fischer, Johann Carl* Froriep, Ludwig Friedrich von Fuchs, Georg Friedrich Christian Göttling, Johann Friedrich August* Graumüller, Johann Christian Friedrich Kastner, Karl Wilhelm Gottlob Lenz, Johan Georg* Nicolai, Ernst Anton Pansner, Johann Heinrich Lorenz Schelling, Friedrich Wilhelm Joseph* Schelver, Franz Joseph* Schenk(e), Johann Heinrich Christoph Scherer, Alexander Nikolaevich Schmid, Carl Christian Erhard Schwabe, Johann Friedrich Heinrich Stark, Johann Christian (I) Stumpf, Johann Georg* Suckow, Laurenz Johann Daniel** Ulrich, Johann August Heinrich** Voigt, Johann Heinrich** Voigt, Friedrich Sigismund

Abbreviations Faculty Dates NFG für Tab 3–5 Philosophy Medicine (1761–1802) OM (B) O (SN) EO

Timespent as lecturer 1790–1802

(1763–1833) – (1779–1847) –

(Fi) (Fr)

Adj., EO –

– PD, EO

1790–1807 1800–1804

(1750–1813

(Fu)

PD, EO

1790–1807

(1753–1809) EM

(G)

EO, HP

1790–1807

(1770–1825) –

(Gr)

PD

1807

(1783–1857) –

(K)

PD

1805

(1745–1832) EM

(L)

1790–1807

(1722–1802) – (1777–1851) AM

(N) (P)

Adj, PD, EO,HP – (PD)

O –

1790–1802 1801–1802

(1775–1854) EM

(Schll)

EO

1798–1803

(1778–1832) – (1732–1798) –

(Sv) (Sk)

EO –

– PD

1803–1806 1790–1798

(1771–1824) OM

(Sch)

PD

1794–1796

(1761–1812) EM

(Schm)

Adj, SN

(1779–1834) –

(Schw)

PD

(1753–1811) –

(St)

SN, EO,O

(1749–1798) – (1722–1801) –

(Stu) (S)

EO O (Physik)

– –

(1746–1813) –

(U)

O (Pol/Mor) –

1798

(1751–1823) EM

(V)

(1781–1850) –

(Vo)

O (Mathe/ Physik) PD

–1790– 1807 1806–1807

EM

1801

1790–1794 1790–1801

Fig. 2. Lectures in sciences in Jena University from 1790–1807. Persons obtaining a chair are marked with two asterisks, associative professors with salary are marked with one asterisk. The abbreviations were used in Figs 3, 4 & 5.

188

O. BREIDBACH

Appendix 2 1790 Chemistry Physics Mineralogy Botany Natural hist. Zoology Geology Geography Meteorology Naturphil. others Σ

1792

1793

1796

1797

S

W

S

1791 W

S

W

S W S

W S

W

S

W

S

W S W

4 2 1 2 2

4 2 1

4 2 1 2 3

3 2 1 1

1 3 1 3 4

2 2 1

1 3 1 3 3

5 3 1

4 3 2

5 4 2 2 2

4 2 3

1 2 2 3 2 1

1 2 1

1 1

2

1

1

3 3 1 3 1 1 1 1

12

7

14

1

1 1

1

1795 4 3 2 2 2

1 2

1798 1 2 2 3 2

1 1 2

1 3 1

1 1

1

1

12

1 13

9 21

1799 Chemistry Physics Mineralogy Botany Natural Hist. Zoology Geology Geography Meteorology Naturphil. others Σ

1 3 1

1794

9

13 7 20

22

1800

19

11 13 9 25 22

10

1 8 18

1801

1802

1804

1805

1806

1807

S

W S

W

S

W

S W S

W

S

W

S

W

S

W

S

W

Σ

1 3 2 1 2

1 2 1

1 3 1

1 1 1 2 1

1 1 2

1 1 1

2 1 1 2

1 2 1 1 1

1 2 1

3 1 2 3 2

1 2 1 2

2 1 1 3 2

1 1 2 1

1

2 1 2 2 1

2 1 1 1 2 1

1

1

1

1

1

1

2 1 2 2 2 1 1

6

9 4 13

1 1 7 12 7 13 19

74 73 49 48 42 15 14 9 2 2 6

10

6

1 2 1 3 3

1 2

2

2 1 9 8 12 7 17 19

1803

16 10 11 8 26 19

1 1 1 1 1 1

1

1 1

1

7 13

6 7 13

6

16

8

1 11 19 334

Fig. 3. Number of various disciplines announced under category of Naturwissenschaften in Jena University; S: Summer: Winter. (Adopted from Bach and Breidbach 2001)

addition to theoretical and experimental physics from 1792 to 1799). In all that time, between 1790 and 1807, there were only five lectures on electricity and magnetism. Thus, 80% of the courses were delivered in experimental physics. In addition the technician (Hofmechanicus) Schmidt gave courses on mechanics (Ziche 2001; 232). Voigt and Suckow gave a lecture on mechanics and optics, Suckow gave one lecture on the theory of heat, and Pansner, one on acoustics. The five lectures on electricity were: the 1st a seminar on electrical experiments delivered by the technician Schmidt in 1788. The 2nd on the theory of celestial electricity was given by the mineralogist Lenz in the same year. In 1796

S

1790

elektr. Versuche Th. Exp. Ph. S+V Mechanik u. Optik phys. Exp. Exp.phys. Th. u. prakt. Exp. physik 1.Teil der allg. Naturlehre Elektr. u. Magnetismus Bew. d. Körp. Th. Physik

Appendix 3

S

S+V S+V

W

1791

S+V

W

V S+V

S

1792

S+V

W

S+V+Fi

S

1793

S+V+Fi

W

W

S

1795

Fi

V S+V S+V S+V

S

1794

S+V

W

S+V+Fi

S

1796

S+V+Fi

W

S+V+Fi

S

1797

S+V+Fi

W

Fi

S+V+Fi

W

(continued)

S+V

S

1798

THE CULTURE OF SCIENCE AND EXPERIMENTS 189

Fi

S+V

S

1799

S

S+V S+V

W

1800 S

S

S+V V

W

1801

V

W

V

S

1802

V

W

V

S

1803

V

W

Fi

V

S

1804

Fi

V

W V

S

1805 S

V+Fi V

W

1806

V

W V

S

1807

V

W

Fig. 4. Lectures given in physics in Jena University from 1790–1807, the title of the lectures are given in German, lecturers were: Fi = Fischer; P = Pansner; S = Suckow; Schm = Schmid; U = Ulrich; W = Wiedeburg. (Adopted from Bach and Breidbach 2001)

elektr. Versuche Th. Exp. Ph. Mechanik u. Optik phys. Exp. Exp.phys. 1.Teil d. allg. Naturlehre Th. u. prakt. Exp. physik Elektr. u. Magnetismus Bew. d. Körp. Th. Physik

(continued)

190 O. BREIDBACH

B

Syst. Chemie Stoffe d. ch. Körper, d.d. Min. Populäre Ch. Med. Exp.ch. (+ mat. med.) Öcon. Ch. F Chem. Convers. Von d. Salzen F Gesch. d. Ch. Technologie Zerleg. /Lösung d. G Chemiekal. Chemie G Im gem. Leben vork. chem. Erscheinungen Antiphlog. Ch. Litteratur u. Nom. d. neu. Chemie Gaschemie Ch. Analyse Physiol. d. anorg. Körper Dünger

1791

F

G

G

N

B

S

F

B

W

1790

S

Appendix 4

G

B

B

W

1792

G

S

G

B

W

1793

G

S

G

W

1794

G Sch

Sch

S

1795

Sch

S

Sch

W

1796

F

Sch

S

Sch

F

Sch

W

S

1797

Sch

Sch Sch Sch

G+Sch G+Sch G+Sch G+Sch G+Sch G Sch

S

Sch

W

G

1798

G

W

(continued)

G

W S

THE CULTURE OF SCIENCE AND EXPERIMENTS 191

S

1799

G

W

G

S

1800

G

W

G

S

1801

G

W

G

G

S

1802

G

W

G

S

1803

G

G

W

G

S

1804

G

W

K

G+K G

K

WS S

1805

G

G

S

1806

G

W

Fi

G

S

1807

G

G

W

Fig. 5. Lectures given in chemistry in Jena University from 1790–1807, the title of the lectures are given in German, lecturers were: B = Batsch; F = Fuchs; Fi = Fischer; G = Göttling; K = Kastner; S = Suckow; Sch = Scherer. (Adopted from Bach and Breidbach 2001)

Syst. Chemie Stoffe d. ch. Körper, d.d. Min. Populäre Ch. Med. Exp.ch. (+ mat. med.) Öcon. Ch. Chem. Convers. Von d. Salzen Gesch. d. Ch. Technologie Zerleg./Lösung d. Chemikalien Chemie G Im gem. Leben vork. ch. Erscheinungen Antiphlog. Ch. Litteratur u. Nom. d. neu. Chemie Gaschemie Ch. Analyse Physiol. d. anorg. Körper Dünger

(continued)

192 O. BREIDBACH

THE CULTURE OF SCIENCE AND EXPERIMENTS

193

Suckow gave a lecture on electricity and magnetism, and in 1798 and in 1799 Fischer, likewise, announced two lectures on electricity and magnetism. Thus, the education in physics in Jena University corresponds to the general picture Herz (1989) outlined for an education in physics in the last decade of the 18th century in Germany, describing this as a uniform practice. In Jena this invariant program in physics was found not only in that last decade of the 18th century, but was extended to the first decade of the 19th century. In Jena these physical lectures were also given by those who gave lectures in mathematics like Voigt and Fischer. Thus, in the sciences, the subjects of the lectures followed a certain scheme: In both chemistry and physics, the list of lectures announced presented a common structure throughout the years from 1790 to 1810 without any major change. For physics and chemistry, the courses of lectures were largely unchanged. Chemistry consisted of a general course, accompanied by certain lectures on applied chemistry. In physics the standard lectures were announced as “theoretical and experimental physics.” Thus, there existed the framework for a course of physical lectures. This one was fixed before 1800. The discussions about animal electricity and the philosophical systems that referred to this new branch of science in the late 1790s where not reflected in the university curriculum. 3.2. Physics textbooks used in Jena lectures In Jena University, in the announcements of the lectures the textbooks used were occasionally given. In case of chemistry and physics these were the following: Textbooks used for lectures in physics (courtesy of J. Frercks): Suckow, Laurenz Johann Daniel: Entwurf einer Naturlehre. Jena 17611, 17822 [used by Suckow]. Erxleben, Johann Christian Polykarp: Anfangsgründe der Naturlehre. Göttingen und Gotha 17721, 17772, Göttingen 17843, 17874, 17915, 17946, 3–6. ed. by Georg Christoph Lichtenberg. [used by Suckow, Voigt, and Fischer]. Fischer, Johann Karl: Anfangsgründe der Physik in ihren mathematischen und chemischen Theile nach den neuesten Entdeckungen. Jena 1797 [used by Fischer]. Mayer, Johann Tobias: Anfangsgründe der Naturlehre zum Beruf der Vorlesungen über die Experimental–Physik. Göttingen 18011, 18052, 18123, 18204, 18235 [by Voigt]. According to Clark (1997) in physics these were standard textbooks that were likewise used in other universities. Most important were the textbooks published by scientists from Goettingen. These did not reflect the impact of the new French scientific thinking in the first decades of the 19th century. As Lind (1992) has shown, the new French textbooks were translated into German during the first decades of the 19th century. Thus, in 1804, Hauys Grundlehren der Physik was published in translation by Blumhof in Weimar, but this was not used as a reference in the Jena lectures (Frercks 2003). These French textbooks reflected the new quantified physics, which was adopted even by German textbook authors as

194

O. BREIDBACH

Georg Gottlieb Schmidt and Ernst Gottfried Fischer. Their texts were not used in Jena. The alternative tendency was a perspective that was influenced by the German and particularly Jenensian Naturphilosophie. Kaestner, Siber, Hildebrand, and Weber had published textbooks reflecting this new tendency (Clark 1997). These books were not used as textbooks in physics education at Jena University. Instead, the textbook of Mayer was used (Frercks 2003). Mayer explicitly rejected Schelling’s speculative approaches. The textbooks used in physics education in Jena University do not point to speculative Naturphilosophie as influencial. Rather, they conformed to those of the descriptive physical science as presented by German authors in the second half of the 18th century. 3.3. Jenensian textbooks on physics The use of textbooks from elsewhere was of special interest as there were a number of more or less conventional textbooks written by members of the Jenensian University staff. These are: Physics Suckow, Laurenz Johann Daniel: (17611, 17822) Entwurf einer Naturlehre. Jena Batsch, August Johann Georg Karl: (1791) Versuch einer Historischen Naturlehre, zweyter physikalischer Teil. Halle Fischer, Johann Carl: (1797) Anfangsgründe der Physik. Jena Chemistry Fuchs, Georg Friedrich Christian: (1787) Chemischer Lehrbegriff nach Speilmanns Grundsätzen, ausgearbeitet, und mit neuesten Erfahrungen bereichert. Leipzig Batsch, August Johann Georg Karl: (1789) Erste Grundzüge der systematischen Chemie zum Unterricht für Anfänger und zur leichteren Übersicht tabellarisch vorgetragen. Jena. Scherer, Alexander Nicolaus: (1795) Versuch einer populären Chemie. Mühlhausen Göttling, Johann Friedrich August: (1798–1800) Handbuch der theoretischen und praktischen Chemie. Jena. Döbereiner, Johann Wolfgang: (1811–1812) Lehrbuch der allgemeinen Chemie, zum Gebrauche seiner Vorlesungen entworfen. Jena. –(1816) Grundriß der allgemeinen Chemie, zum Gebrauche seiner Vorlesungen entworfen –(1819) Anfangsgründe der Chemie und Stöchiometrie. Jena Jan Frercks has analysed the physics text books in more detail. According to his results (Frercks 2003), there was neither an influence from the new French trend

THE CULTURE OF SCIENCE AND EXPERIMENTS

195

of a quantification in physics nor the influence of Schellingian Naturphilosophie. Frercks states that Fischer integrates the fundamental of Kant’s work Metaphysische Anfangsgründe der Naturwissenschaft into the introductory chapter of his textbook, however without much impact on the main body of the text. As has been mentioned, this was similarly done by Gren in Halle and even the aforementioned Mayer in Goettingen. The latter was described as an explicit antagonist of Schellingian Naturphilosophie Frercks (2003) has analysed the chapters on electricity in more detail, comparing the books published in Halle, Goettingen, and Jena. According to that analysis, Batsch’s text is a rather descriptive account, which in its argument shows no essential differences to those books used in the lectures in Jena University. Correspondences were shown for the definition of electricity, the classification of electrical phenomena, the description of the equipment; and the phenomena. However, Batsch did not teach physics, and his book was not used by his collegues. In 1793, Voigt published a monograph Versuch einer neuen Theorie des Feuers (An attempt towards a new theory of fire) , which aimed at a discussion of Lavoisier’s theory by referring to an actualised version of the former phlogiston theory. This is no proper textbook, but it possesses an extensive chapter of 137 pages on electricity. This account is more detailed than a textbook contribution, and even tries to provide a theoretical framework. Its argument follows the standard textbook organisation, starting with a classification of electric phenomena, passing on to a description of the electrical machine. It argues about the theoretical understanding of electricity in a criticism of Franklin’s view on the subject, describes the “laws” arrived at, and goes on to the equipment: electric machine, electrophorus, Leyden jar, and condenser. It ends with a description of the organic effects of electricity. Thereby, Voigt describes a whole series of his own experiments. The usual “Goettingen-Halle—type” textbook (Frercks 2003) has left traces in Voigt’s monograph, without, however, making it part of the education. As was the case with Batsch’s book, also Voigt’s monograph was not referred to by the Jenensian physicists in education. Fischer likewise contributed a textbook that—apart from its failures—did not differ substantially from the text of Erxleben used in the Jenensian lectures as reference book. In his history of physics, Fischer is very explicit in his references to Voigt, whom he obviously took as an authority. There were even found some passages on Schelling’s Naturphilosophie and on Ritter’s experiments. Fischer criticised the speculative access to physics Schelling had proposed. He also criticised the philosophical passages of Ritter’s texts, but explicitly referred to Ritter’s experiments. He tried to understand these from the point of view of the experimental practice of Voigt. Until his death in 1823, Voigt himself used Mayer’s text in his lectures. Accordingly, he worked with a textbook with an explicit negative evaluation of the impact of the speculative Naturphilosophie on science. Although conventional in nature, Fischer’s textbook was used by neither Suckow nor Voigt in their lectures. Directly after its publication it was heavily criticised in the Allgemeine Literatur-Zeitung, where a number of failures und misinterpretations were listed. In consequence, it was not well received in the

196

O. BREIDBACH

Jenensian community. Thus, irrespective of the publication of Fischer’s text in 1797, Erxleben’s book remained the basic text for the physical education. Thus, the textbooks used in Jena and the monographs published by the Jenensian authors do not show that the speculative Naturphilosophie influenced their educational practice. In contrast, the texts used and—less strictly—the texts produced were worked out according to a certain standard as set out in the textbook of Erxleben, for example. After all, these texts corresponded to the standard references from that time in German physical sciences. There is neither an effect of the new French physics nor an obvious impact of the philosophical discussions about polarization of nature in these publications.

4. RESEARCH 4.1. Physics What can be said about the publications of these scientists? In regard to that prominent figure in the education of physics, Voigt, who for over 30 years gave most of the lectures in that branch of science and even published his own journal, Frercks has asked whether he can actually be regarded as a physicist. From 1789 to 1797, Voigt published just five original papers, and only two of them dealt with physical problems (in a modern sense). In his Versuch zu einer neuen Theorie des Feuers, Voigt described own experiments. These are designed to specify the various phenomena of electricity, but do not present a theory. Voigt is to be regarded as someone working on various phenomena, but not as someone, being engaged in regular discussion with colleagues (at least in so far as it resulted in journal contributions). The same is true for Fischer, who wrote as well as his textbook a dictionary of physics and an extensive history of physics in eight volumes. There is no original publication by him that dealt with the problem of galvanism. The same is true for Suckow. These men were first of all university teachers. And they taught not only physics but also mathematics, fortification techniques or Kameralwissenschaften. As far as a judgement is possible in regard to the published texts, these people had no great interest to contribute to the new developments in physical science. There is one problem, however, when we try to reconstruct the impact of the discussions on galvanism on the educational programme in physics. The textbooks analysed provide the reference texts for these lectures, but they did not give the actual lectures. Thus, even in the standardised scheme of physical lectures, there might be a change in the actual presentations given to the students, over the various years. The analysis of the publications of the Jenensian lectures gave no hint, however, that there was an interest in that regard. Nevertheless, for a definitive judgement we need access to various lecture notes. We have started a search for such notes, but up to now, we have found only two. A first analysis of the notes of a lecture of

THE CULTURE OF SCIENCE AND EXPERIMENTS

197

Voigt did not reveal any substantial new aspects in our evaluation of his teaching. However, we really need a series of such notes spanning the critical decade around 1800, before such a definite judgement can be made. 4.2. Natural history In natural history, the situation was very different from that in physics and in chemistry before the start of Doebereiner’s career in Jena.4 In the botanical sciences, the situation is different. Batsch published a whole series of monographs and research notes dealing with the natural system in botany. Batsch, as indicated by some new findings of Linné (in prep), followed up his own theory of systematics. The botanical garden he established after 1797 was designed to present this new natural system. Here was a very active and well received scientist. His interests, however, were directed not towards experimental and analytical sciences, but natural history. Thus the situation differed from that described in more detail for physics. Batsch was following up his own scientific thinking and concepts (Jahn 1963). He was in contact with Goethe and shared Goethe’s approach to a typology of nature. His successor as professor in botany and as director of the botanical garden, Schelver, was well known as a devoted follower of the speculative tendencies of natural philosophy (Müller 1992). This he was, even before he was engaged for Jena. Thomas Bach (2001a) has found lecture notes of a course Schelver delivered at the University of Halle which is explicitly designed as a course in Naturphilosophie following Schelling’s ideas. When Schelver left Jena to take up an appointment in Heidelberg University his successor was Voigt. Voigt explicitly worked in the conceptual frame of Goethe’s idea of a metamorphosis of nature. In botany, Voigt published a textbook that tried to establish both a new terminology and a new methodological approach (Voigt 1808). That approach was little more than a direct transfer of Goethe’s idea of the metamorphosis of plants into a new system of botany designed according to Goethe’s principles. In 1807, Lorenz Oken was engaged in the medical faculty to represent comparative anatomy. Oken was follower and friend of Schelling. He not only wrote his own Naturphilosophie, but worked out a new systematics for natural history adapting Schellingian principles for biological classification (Bach 2001b; Breidbach and Ghiselin 2003). Oken was a very active scientist. In his later days he became the first president of Zurich University (Schäffner 2001). 4.3. Romantic physics? Apart from Oken and Schelver, the persons who formed the academic staff of the natural sciences in Jena were not connected with the romantic sciences in the recent research literature (for surveys see: von Engelhardt 1997; Knight 1998; Poggi 2000). And—as far as our analysis goes—there is indeed no close 4

For situation in chemistry cf. Döbling 1928; Chemnitius 1929.

198

O. BREIDBACH

connection between the university physics lecturers and the ideas of the speculative Naturphilosophie that—according to literature—gave rise to a romantic science in Jena. This literature, however, is concerned with Johann Wilhelm Ritter who in 1798 published his monograph Beweis das ein beständiger Galvanismus den Lebensprozeß im Tierreich begleitet (Richter 1997). Between 1800 and 1805, he wrote his Beyträge zur näheren Kenntniß des Galvanismus. As has been mentioned, Ritter was not a member of the academic staff of Jena University though—as been mentioned before—he had permission to deliver one course on his subject, animal galvanism. This course was never announced in the official lists of lectures printed by the university. Ritter, thus, cannot be regarded as part of the university department of physics. Nor was he offered laboratory space. He gave presentations of his ideas and experiments on various occasion in the Society of Natural Sciences (Gesellschaft für Naturforschung), which at that time, was not part of the university. The main part of its membership in Jena, however, were students of the University (Ziche 2001). Laboratory equipment was offered to Ritter directly by the government of Sachsen-Weimar-Eisenach (Richter 1988), he could build up his column with that money, and brought people like Humboldt, Goethe, and Frommann to his flat; yet that was done alongside rather than within the university. His research nevertheless was received by Jena University teachers. Fischer mentioned it in his history of physics. Fischer discussed the work of Ritter, as the work of someone in the experimental tradition of Voigt. Thus Fischer was concerned with Ritter’s experiments though he did not argue about Ritter’s theories. Accordingly, Ritter’s research was part of the discourse about physics in Jena. Such discussions, however, were not part of the curriculum of university studies. Ritter gave talks in the Naturforschende Gesellschaft (Richter 1997).

5. MEDICAL EDUCATION AND PRAXIS 5.1. Overview In the medical faculty, the situation of scientific education is less clear than in the philosophical faculty, which the sciences were part of. The main figures around 1800 were the physicians Hufeland and Stark, and the anatomist Loder. Hufeland was standing in the tradition of 18th century anthropology (Giese and Hagen 1958; Genschorek 1977). His famous work über die Kunst das menschliche Leben zu verlängern (The art of prolonging human life) is a practical treatment without any theoretical reflection. He was the editor of a famous journal on practical medicine and the author of many monographs on questions of practical medicine. Stark likewise was engaged as a physician and taught about practical medicine. Loder was an anatomist who made his fame with a thorough treatment of human anatomy, published in his Tabulae anatomicae from 1794–1803. Apart from these leading figures, there was a whole series of lectures that—with the exception of the aforementioned Lorenz

THE CULTURE OF SCIENCE AND EXPERIMENTS

199

Oken—have not been treated in much detail until now (Giese and Hagen 1958). Within the lectures offered in Jena University around 1800, one can find special lectures concerned with electrotherapy that are of interest in the context of the present paper.

1. Chemistry 2. Surgery Med. Practice Anatomy Ackermann, J. F. Batsch, A. Joh. G. Bernstein, J. G. Bretschneider, F. F. Croneberg, A. J. Froriep, L. Fr. Fuchs, G. Fr. C.

1804–(1805)

Fuchs, J. Fr. Gruner, Chr. G. 1804–(1815) Hallbauer, Fr. J. Hardege, H. Z. v. Hellfeld, C. A. F. Hellfeld, J. A. Hufeland, Fr. Hufeland, W. Fr. Kieser, D. G. Kilian; K. J. Löbenstein-Löbel, E. Loder, J. C. Martens, F. H. Neubauer, J. E. Nicolai, E. A. 1775–(1802) Oken, L. Renner, Th. Schelver, Fr. J. Stark, C. W. Stark, d. Ä., J. C. 1803

1805

Stark, d. J., J. C. Succow, W. C. Fr.

Voigt, Fr. S. Walch, Fr. A.

1816

3.Theory Botany

PD

Med. Associate Professor

SN

1787–(1802) 1802–1803 1800 1801–(1802) 1812–1815 1801–1803 1803–1804 1783–(1801) 1783–1793, 1804– (1813) 1801, (1804) (1804) 1776–1803

1779–1783

1775–1793 (1806) 1783–1819

1811–1812

1812

1775 1792–1801 1812–1818 1801–(1802) 1811–(1818) 1778–(1803) 1804–(1805) 1775–1777

1804– (1811) 1805–1819 1811, 1819 1793–1794 1795–1801

1807–1812 1816–1819 1803–(1806) 1816–1819 1779–1782 1783–1802

1796–1804 1801–(1802)– (1805), 1810 1810–1818 1805– (1817) 1817–1819 (1818)

Fig. 6. List of lecturers in medicine between 1790 and 1820. (Courtesy of K. Regenspurger)

200

O. BREIDBACH

Faculty

Semester

Lecture

Lecturers

Medicine

SS 1797

Succow

Mathematics/ Physics

WS 1786/87

Naturwissenschaften

SS 1796

Philosophy

WS 1788/89

Philosophy

SS 1783

Der medicinische Gebrauch der Electricität Die Electricität und der Magnetismus nach eigenen Dictaten Die Theorie der Electricität und des Magnetismus Die Theorie der himmlischen Electricität Die Mechanik, Architektur, Gnomonik und Electricität praktisch

Batsch

Succow Lenz Hofmechanikus Schmidt

Fig. 7. List of lectures concerned with electrotherapy, animal galvanism, and electricity. 5.2. Franz Heinrich Martens The lecturer Franz Heinrich Martens wrote a monograph entitled: Complete manual for the therapeutical application of Galvanism. It was printed 1803 in Weißenfels and Leipzig, and became a standard text in that branch of sciences. There is one particular item given already in his subtitle: Together with a history of that remedy in regard from the primary origin of this branch of medicine to the newest application. In fact, Martens set galvanism into the tradition of electrotherapy that—for example—had been provided already in the 1750s by Johann Heinrich Winkler, one of the teachers of Goethe, in Leipzig (Winkler 1747). Martens’ work added to the application of the voltaic pile to electrotherapies. He described a theoretical framework for electrotherapy, and discussed the then new findings in animal galvanism. Hereby, Martens explicitly referred to Ritter. Ritter’s proof that a continuous galvanism is found alongside the process of life in the realm of animals—Martens argued—might not be correct in each part of the argument. Furthermore, parts of the argument just adopt a new terminology to old findings. Nevertheless, a new type of research may in the end give new insights of practical value for medical treatment and in a long run will strenghten the importance of electrotherapy in practical medicine.5 Martens did not set Ritter’s research into a speculative framework but treated it as part of a tradition in electrotherapy. Medical treatment was greatly facilitated by the new instruments developed by scientists interested in galvanism and electricity—especially by the Voltaic pile which allowed much easier and cheaper treatment. Yet even more important was

5

“Wenn nun aber auch Ritter’s Beweis, daß ein beständiger Galvanismus den Lebensprozeß im Thierreiche begleite, nicht so ganz bündig seyn sollten oder schon bekannte Sachen in diesen Worten nur mit einem neuen Namen belegt vorgetragen wären, so läßt sich doch mit Recht erwarten, daß durch die immer größere Vervollkommung der angegebenen Wissenschaften die Theorie des Galvanismus mehr Bestimmtheit und die Praxis eine dem menschlichen Organismus gemäßere und ausgedehntere Anwendung erlangen würde” (Martens 1803: 34).

THE CULTURE OF SCIENCE AND EXPERIMENTS

201

the detection of an animal electricity or an animal galvanism. According to this finding, electricity was not an artefact to be used in medical treatment, but an essential physiological characteristic of man. Thus, Ritter’s proof of a coexistence of galvanisms and the process of life apart from the details of Ritter’s argument was of utmost importance to Martens. He just had to point to that statement to declare that electrotherapy was a central part of applied physiology. Martens did not go into the details of Ritter’s proof. He referred only to the general statement as he regarded Ritter as an excellent observer, citing him eight times in his book. Ritter was thus treated as someone who had added important new findings to the tradition of electrotherapy. For Martens this justified positioning his medical discipline as a core discipline in his faculty.6 The aim of Martens’ monograph, thus, was a descriptive one.7 Thereby Martens explicitly described the theory of Volta and the ideas on Galvanism presented by Tittmann. Tittmann considered galvanism to be an epiphenomenon. He discusses various chemical reactions that were induced by electricity and thought galvanism to be such an effect of chemical reactions comparable to the production of heat in various of such reactions. Martens’ conclusion on such divergent theories is explicit: One has to be descriptive to show the various useful effects of electrotherapy. And one has to omit theoretical discussions from a manual in electrotherapy. This he did even in regard to Ritter’s work. In two short passages Martens’ referred to Humboldt (Martens 1803: 86, 94) as he did with Ritter’s experimental results. Martens did not deal in any detail with Ritter’s theoretical background, however. In so far, Martens’ approach to galvanism and electricity came pretty close to that of Fischer, who in his definition of electricity gave a solely descriptive definition of the phenomenon, explicitly abstaining from theory.

6

7

“Ritter erklärt diese Erscheinung ( attraction and repulsion of galvanic elements) sehr scharfsinning nach dem Gesetze der elektrischen Mittheilung und Vertheilung. Diese Erscheinungen sind für die genaue Kenntniß des Galvanismus zu wichtig als dass sie nicht auch den praktischen Arzt interessieren sollten, und dadurch glaube ich es entschuldigen zu können, Daß ich dieselben hier anführe” (Martens 1803: 68). (Ritter explained these phenomena according to the law of electrical exchange and electrical dispersion. These phenomena are of utmost importance for an understanding of galvanism. Accordingly even a physician should know them. Thus, I thought it necessary to explain these in my present text.) “Zu dieser Untersuchung ist es nothwendig, daß ich mich auf den Begriff des Galvanismus, auf die Entwicklung und Verbreitungsart desselben, so wie auf die sogenannten Leiter und Nichtleiter derselben beziehe; daß ich ferner die beym Galvanismus vorkommenden Erscheinungen, die Oxydation , die Zersetzung des Wassers, die Funken, die Erschütterung oder Schläge, die Anziehungen u.s.w. beschreibe und darstelle, den Einfluß der Luft und die Temperatur derselben auf den Galvanismus erweise, die wichtigsten und allgemeinsten Erscheinungen mitteile, welche sich bey der Einwirkung desselben auf den thierischen Organismus und auf die einzelnen Theile und Systeme desselben äußern , die neueste und wahrscheinlichste Theorie desselben mitteile und das Verhältnis des Galvanismus zur Elektrizität nach allen Gründen prüfe und bestimme…” (However, to enfold such a descriptive study in that area, it is necessary that I explain the term galvanism, its use, and its application. I have to explain conductors and isolators, furthermore I have to describe the phenomena accompanying galvanism like oxidation, hydrolysis, lightning, shock, attraction, and so on. I have to describe the influences of air, of temperature on galvanism. I have to report the most important and most general phenomena, elicited by galvanism on organisms and organic tissues. Thereby I have to report on the newest and most probable theories. Finally, I have to discuss the relation of galvanism to electricity…).

202

O. BREIDBACH 6. PHENOMENOLOGICAL SCIENCE—FISCHER’S TREATMENT OF ELECTRICITY

Fischer in his definition of electricity printed in his encyclopaedia gave a complete list of phenomena (Fischer 1798: 861f): Electrizität heißt derjenige Zustand eines Körpers, worin der letztere Körper anfänglich anzieht, nachher wieder zurückstößt, wenn sie ihm hinlänglich genähert werden, mit einigen ihm nahe gebrachten Körpern, z.B. mit dem Knöchel oder der Spitze des Fingers einen stechenden und knisternden Funken gibt, einen gewissen süßlichen Geruch, der nach Urinphosphat richt, um sich her verbreitet, gewisse anderen Körper ebenfalls die Eigenschaft mittheilet, eben diese Wirkungen hervorzubringen und dergleichen bald anzuführende Erscheinungen mehr. Oftmals versteht man auch unter dem Worte Elektrizität nicht allein diesen beschriebenen Zustand des Körpers sondern die Ursache selbst, welche diese Wirkung hervorbringt. In dieser Bedeutung soll aber hier die Elektrizität nicht genommen werden. (Electricity is that state of matter, where the latter is attractive to another body at first, but repulsive at second, when the two bodies have been put close together. It elicits a small painful spark, when a knuckle or a finger comes pretty close. The smell elicited is slightly sweet like urine phosphate. It electrifies certain other bodies and so on. Often the term electricity is not used to define such phenomena, but to speak of the cause eliciting such effects. Electricity should not be taken in that regard, here, however.) Electricity seems to be defined simply as that which is being produced by electrical machinery. Fischer—as he did in his history of the physical approach to electricity—stuck to a description. What he was saying, was the story of different variations of experiments done in course of the 18th century. He described how different experimental settings had been handled in the course of that century outlining the development of electrophysics in the respective experimental procedures. Fischer referred to Ritter, describing how far the latter had varied former experimental settings and what the results of these were. That reception is important from two directions. Firstly it shows us, how Ritter’s speculative attempts were received in Jena. Ritter’s actual experiments, not his theories were described. That was explicit in the positive reception of Ritter and Humboldt by the Italian community (Breidbach and Frigo 2003). Both were seen as providing important support to the details provided by Volta, Galvani, Aldini, and the others. Particular interest was devoted to Ritter’s self-experiments, which were described as standing in the tradition of good experimental practice known from natural history and the sciences from 18th century German science.8 Fischer and Martens show a comparable reception. There is no trace of a speculative infection in these authors. On the contrary, Fischer explicitly argued against the use of Schellingian principles in physics in the first volume of his history of physics. That shows us that a specific romantic style of sciences did not develop of the sciences at Jena University around 1800. 8

For reception in France, cf. Klengel 2001.

THE CULTURE OF SCIENCE AND EXPERIMENTS

203

But there is a second observation to be made from Fischer’s text. Fischer describes his science as a gradual extension of experimental practices. He argues against deductive preconceptions and explicitly limits himself to the description of single observations. Thus, in that branch of physics where no theory was well established, people like him relied on induction. His science was data collection, and it had been this already in 1790, before the onset of Schelling’s philosophy. So far his account on the physics of electricity seemed to anticipate Schleiden’s later programmatic view of science as an inductive (and not just empiric) procedure (Charpa 1989).

7. EXPERIMENTAL PRACTICE 7.1. Reconstruction of an electric machine The question, however, is how far the description in such texts actually reflected experimental practice. Texts were used as a reference or as science manuals. Thus, Martens explicitly wrote that he designed one of his experiments according to the procedure described in Voigt’s magazines (Martens 1803; 118). In my group, we have asked, what the standards of published experimental protocols were in Jena around 1800. The aim of this approach is to understand how far one can actually separate description and theory in the various texts on electricity at that time. A first step in solving that problem was the reconstruction of the electricity machine manufactured by Schmidt, a technician engaged at the court of Weimar, who also gave courses at Jena University (Weber 2002). He wrote out a detailed paper on the construction of that machinery (Schmidt 1773). This allowed us to do the reconstruction—even remanufacturing the material the machinery had originally been made of. This reconstruction gave us insights into the state of arts in mechanical profession and in the production of glass in Weimar Jena about 1800, showing a highly developed technical practice in that region (Weber et al. 2003). We investigated how far the machinery could be constructed just following Schmidt’s description. The observation was that the text of Schmidt did not allow us to do this. It was possible only with reference to the illustrations. Schmidt had added to his publication. These allowed us to calculate the relative dimensions of the machinery, compare these with some measures Schmidt had outlined in his text and, thus, to redo the complete construction. Luckily, parts of that machinery have been preserved in the Goethe-Museum in Weimar, so we could easily test wether our reconstruction was correct. The pictures had not been to scale. However, by comparing it with descriptions given in the text, one could achieve in an adequate scaling—compared to the original preserved in Weimar. But apart from that problem, the outcome was that only by the paper of Schmidt alone, one would not succeed in a proper reconstruction. But if one was able to reconstruct the state of technologies especially in regard to glass

204

O. BREIDBACH

Fig. 8. Engraving of Schmidt (1773). production practices at that time, one could at least approximately succeed in a construction of such a machine.9 That meant that the variations tolerated in the manufacture were not discussed in the publications. The scientists referred to an inexplicit knowledge being present in the traditions of manufacturing. The detailed reconstruction of the failures in the reproduction of the machinery allowed us to identify these, and thus, to determine how far they were part 9

For the problem of an evaluation of “success” in experimental history of science see: Sichau (2000).

THE CULTURE OF SCIENCE AND EXPERIMENTS

205

Fig. 9. Photography of the glassworks done in our effort to rebuild the electric machine of Schmidt.

of an explicit laboratory culture or were just referring to techniques employed more generally. In regard to the electrical machinery where we could point to a highly specified manufacturing tradition, these machines were not part of an explicit laboratory culture found exclusively in the physical sciences. They were manufactured using the current high standards, especially in glass production (Weber et al. 2003). No special techniques were required in the production of such experimental devices. Thus, the inexplicit knowledge of the former readers of such a text was nothing more than a reference to the common technical standards of the day. Thus Martens, in his attempt to do his experiments according to the descriptions in Voigt’s magazine, demonstrates the existence of such common standards. 7.2. Experimental culture Any question as to the specificity of an experimental culture had to ask to what extent such broader characteristics were made explicit. Accordingly, I propose a definition of an experimental culture. An experimental culture is seen not just as a technique employing a certain machine but as a technique that employs a certain machinery or a way to handle something in a certain way. That way is reflected by a theory employed for an interpretation of those things that one should register and the order criteria by which such a registration is compared to other observation within a certain branch of science. If a theory referred to something done in a particular way, the theory was developed besides the experiments. Only if the theory changed the standard would a separate experimental culture have been created. The Jenensian reception of Ritter tells us that this was not the case with him. Ritter’s experiments were regarded as parts of an experimental

206

O. BREIDBACH

tradition common to scientists of his day. In the experiments, the various effects produced were registered and added on to the sum of knowledge. As there was no theory to explain the inherent structure of the phenomena tested, there was no possibility to sort out essential and accidental parts of an experimental set up or an experimental procedure. This can be demonstrated in both the electrical machinery and in the descriptions of the experiments. From our catalogue of various electrical machines found in 18th-century central Germany, it can be demonstrated how various parts of the machinery persisted and what was changed in the construction throughout the 18th century (Weber 2002). This allows us to describe what was seen as being essential and what was not. Enlisting such characteristics, one could even develop a kind of phylogenesis of the various types of that machine, and thus, identify those parts of the machine that were seen as essential in the course of the 18th century and those that were not. Some of the more ornamental particulars, like the leather skirts around certain parts of the machinery, can be demonstrated as the most persistent parts of that construction (Hackman 1978; Weber and Breidbach, in prep).

Fig. 10a. Electric machine of Johann Gottlob Krüger, pictured 1745. (Courtesy of H. Weber)

THE CULTURE OF SCIENCE AND EXPERIMENTS

207

Fig. 10b. Electric machine of Carl Gottlob Kühn pictured 1788. (Courtesy of H. Weber)

8. NATURPHILOSOPHIE 8.1. Schelling Jena then, seemed to offer something not too different from the general tendencies in the sciences of that time, and thus, Jenensian scientists were received well in various different places. The situation in the philosophy of nature was different, however. In Jena around 1800 Schelling proposed a concept of “pure science” (eine Naturwissenschaft im strengsten Sinne des Worts) (Schelling 1799: 275) to replace the former descriptive natural history. His speculative Naturphilosophie

208 S 1797 W W S W S

1797 1798 1799 1799 1800

W S W S

1800 1801 1801 1802

W 1802 S 1803

W 1803 S 1804

W 1804 S 1805

W 1805 S 1806 W 1806

S 1807 W 1807 S 1808 W 1808

O. BREIDBACH Schmid: Zoonomie oder Philosophie der organischen und anorganischen Natur Schmid: Zoonomie Schelling: Philosophie der Natur Schelling: Philosophie der Natur Schelling: Organische Naturlehre Fischer: Philosophie der Natur Schmid: Physiologie oder Philosophie der organischen und animalischen Natur Schelling: Philosophie der Natur Schelling: System der gesamten Philosophie Schelling: Allgemeines System der Philosophie Schelling: Die gesamte spekulative Philosophie Schad: Absolute Harmonie der Natur mit der Ichheit Schelling: Allgemeines System der spekulativen Philosophie Schad: Naturphilosophie nebst Transcendentalphilosophie Schad: System der Natur- und Transcendentalphilosophie Krause: Naturphilosophie Hegel: Enzyklopädie der Natur Schad: Naturphilosophie Krause: System der Natur- und Transcendentalphilosophie Krause: Elemente der Naturphilosophie Gruber: Naturphilosophie Henrici: Naturphilosophie Schelver: Naturphilosophie. Ein System der Naturwissenschaft enthaltend (a) Naturphilosophie, (b) Elementar-Physik, (c)Geognosie, (d) Physiologie des Organismus oder Phytonomie, Zoonomie und Anthropologie Hegel: Die ganze Wissenschaft der Philosophie Krause: Ein System der allgemeinen Philosophie Henrici: Naturphilosophie mit Rücksicht auf den medicinischen Teil dieser Wissenschaft Hegel: Das ganze System der praktischen und theoretischen Philosophie Hegel: Realphilosophie Henrici: Philosophie der Natur mit Rücksicht auf Medizin Hegel: Philosophie der Natur und der menschlichen Vernunft Henrici: Philosophie der Natur mit Rücksicht auf Galls Schädellehre Hegel: Philosophie der Natur und des Geistes Schelver: Philosophie der Natur Henrici: Die Philosophie der Natur Hegel: Philosophie der Natur und des menschlichen Verstandes Henrici: Naturphilosophie Oken: Naturphilosophie Oken: Philosophie der Natur Oken: System der Naturphilosophie

THE CULTURE OF SCIENCE AND EXPERIMENTS S 1809 W S W W

1809 1810 1810 1811

W W W S

1812 1813 1814 1815

W 1815

209

Oken: System der Naturphilosophie Henrici: System der Naturphilosophie Oken: System der Naturphilosophie Oken: System der Naturphilosophie Oken: Naturphilosophie Oken: Naturphilosophie Oken: Höhere Gegenstände der Naturphilosophie für Eingeweihte Oken: Philosophie Oken: Naturphilosophie Oken: Naturphilosophie Bachmann: System der Philosophie (Teil 2 = Physik oder Naturphilosophie) Oken: Naturphilosophie

Fig. 11. List of lectures on Naturphilosophie in Jena from 1797–1815. (Adopted from Breidbach 2000 and Bach 2001a)

was directly relevant to the sciences of his day, which tended to become lost in a heap of unstructured data (Breidbach 1982). Schelling’s concept of a new science of nature tried to envision science not simply as an attempt to describe particular natural objects but as an effort to analyse the principal qualities of nature (Neuser 1995, 1997). This was based on the idea that it is possible to characterize fundamental reactive properties of nature. Such properties were believed at that time to be isolated in the reactive principles that underlie the different manifestations of electricity (Breidbach 1982). Consequently the concept of polarity became attractive, for it seemed—at least for some years—to describe the principal manifestations of reaction modes in nature (Moiso 1994). Schelling’s idea was that nature could be decoded by analysing such a principal dynamic organization. Thereby he thought he could offer something like a world formula (Weltformel) to explain nature without referring to God or any other external principle (Engelhardt 1997). 8.2. Schad, Krause, Schelver, and Oken That view was well-received by group of younger philosophers in Jena. These however challenged the original Schellingian programme (Breidbach 2000; Bach 2001a). For them, it was not the situation in the sciences, but the position achieved by Schelling, when dealing with the sciences, which was of value. They took the principles Schelling had decoded from his analysis of the situation in the sciences as starting points for their own philosophies. This meant their arguments were founded on purely theoretical statements. They ripped off their philosophies from a dialogue with the sciences of the day.

210

O. BREIDBACH

After 1800, nevertheless, it seemed to be opportune to lecture on the philosophy of nature according to Schelling, like Krause, Schelver, and Schad. The question is, to what degree, these scientists adopted the Schellingian view. As I cannot go into details here, I will just, ketch some general notions (for details see: Breidbach 2000; Bach and Breidbach 2001). All these philosophers started their approach with Schelling’s principles. Doing so, they perverted Schelling’s concept of a pure science. All of them produced booklets presenting their specific variation of Schellingian Naturphilosophie, and explicitly referred to Schelling as the one that had secured or discovered the world formula. What they did, thereafter, was simply to apply it. This meant that all these philosophers started with schemes that were no longer tested, and any discussion of them is lacking in their texts. Even the mathematical philosophy of Krause was nothing but an “ars combinatoria sensu Schelling” (Krause 1804). Thus, in contrast to Schelling, these philosophers omitted any explicit reference to scientific problems. In their texts, empirical statements were introduced only as illustrations of deduced principles. Consequently, they did not discriminate between scientific and everyday experiences. In addition, in some of them a messianic impetus to remission in the sciences had developed: One of them, Schad, even tried to offer a lecture on electrical phenomena in the sciences at Jena University. He evidently intended to do his demonstrations in a deductive way, but was stopped by Schelling and Hegel from doing so. That kind of philosophy set itself apart from the sciences. The ideas Schelling had in mind, to give the sciences a theoretical framework in which something could be falsified was not relevant for that generation of scientists. That set themselves apart from sciences and, thus, perverted the original Schellingian approach. Lorenz Oken was the only philosopher who actually tried to systematize natural history using Schellingian principles (Bach 2001b). Thus, his attempt can be regarded as an approach to verify the validity of Naturphilosophie. How far that was done, cannot be part of this paper. As has been mentioned, in his time, Oken was quite successful in doing so. He was well-received in the sciences, and had a great influence in the formation of a nationwide dialogue within German sciences via the Versammlung deutscher Naturforscher und Ärzte (Sudhoff 1922). As has been mentioned, he was also well-received in the English biological tradition (Breidbach and Ghiselin 2003). In that respect, Schellingian philosophy was not outdated in the 1870s. Oken is just one—Jensensian—example for this. Thus, when arguing against Schellingian philosophy, the Jenensian philosopher Fries was arguing against an influential concept (Breidbach 1999c). He created an explicit antithesis against that dominant tradition so well-received in Berlin in his days. 8.3. Schleiden The botanist Schleiden, famous as one of the founding fathers of cell theory in biology—via Apelt—was a follower of Fries’ philosophical approach (Glasmachern 1989). When Schleiden referred to Kant, he explicitly referred to Kant as seen by Fries (Breidbach 1998a). With him, and—comparably—with Justus von Liebig and Johannes Müller, the whole attitude of an anti-speculative but Kantian approach

THE CULTURE OF SCIENCE AND EXPERIMENTS

211

was in fact a remake of Fries, and thus a reductive perspective on that pre-idealistic concept (compare von Engelhardt 1992). Nevertheless, Schleiden was an eminent figure in the theoretical fight of German sciences against the speculative tradition that was dominant until the 1840s in German science (Breidbach 1998b). Around 1840, figures like Oken and Nees von Esenbeck and their concepts of natural history were well established in sciences (Bohley 2001). The new sciences following an analytical and inductive approach had to fight against these figures. In the 1840s, when Schleiden (1844) started that fight, however, he did not directly attack their positions but looked back on the beginning of the tradition these figures were situated in. Schleiden argued not with his contemporaries but with the Jenensian positions from 1800. Thus, in his criticism he established a view of a specific Jenensian tradition. It should be sorted out how far that perspective was actually instrumental in creating an idea of a Jenensian community in science (effective up to now) that in fact, as has been demonstrated here, did not exist. 8.4. Conclusion In spite of the proclamation of an analytical science working with inductive methods (Schleiden 1844), until the second half of the 19th century science in general and biology in particular was dominated by speculative approaches such as those put forward in the romantic sciences (Cunningham et al. 1990). On that basis it has been shown that romantic science was somewhat of a general European trend in the analysis of nature (Knight 1984; Poggi 2000). Nevertheless, research traditions focused on a specific German tradition of romantic sciences, particularly connected with the Schellingian Naturphilosophie can be located prominently in Jena University (Breidbach 2000). These however, are found in the philosophical department sensu strictu. The natural and the medical science did not develop such a specific Jenensian tradition. In those branches of science a well organised standard of lectures and experimental practice were the order of the day. What can be shown, however, is, that the idea of a post-romantic birth of inductive sciences has to be corrected. Induction was present before 1800. Speculative Naturphilosophie does not interfere with the inductive sciences, but initially tried to find a structure for those sciences that tend to lose their systematics in masses of detail. That young group of philosophers who tried to praise Schelling as the messiah of a new philosophic religion like Schad or Schelver were just an episode rejected already before 1810 (Müller 1992; Bach 2001b). Oken’s attempt was a different one. Oken secured at least some of the perspectives of the original Schellingian approach and was effective in that both in science politics and in natural history until the middle of the 19th century.

9. MICROCOSM JENA Apart from a number of details, our micro-historiographic approach has outlined the complexity of the interactions we have to study in a reconstruction of the culture of science in Jena around 1800 (Breidbach and Ziche 2001). At that time

212

O. BREIDBACH

science is reorganising its ordering categories for both the registration of their objects of study and the organisation of scientific studies. Thereby, philosophy and science interact, but such interactions are found only within certain persons. Thus, there is not simply one network of researchers one has to study but a complicated hierarchy of networks not only employing scientists but also students, politicians, and technicians. Ritter when doing his research in Jena was a student. Goethe held a position associated with the government in Weimar that allowed him to initiate certain research programmes (Schmid 1979; Steiger 1986; Müller 2001). Thereby, science was part of a political culture. It would be insufficient, however, to extend an analysis of scientific discourses only in that dimension. The development of sciences was also embedded in certain technologies. Studies of electrical phenomena were possible only when instruments were provided and techniques were available to build special experimental equipment. Thus, the microcosm analysed is not a separated and autonomous cosmos of scientific ideas but is part of a cultural practice that recruits a wide range of interactions from techniques to aesthetics.

ACKNOWLEDGMENTS This study summarises the results obtained in the first years of the project on speculation and empirical knowledge within the Sonderforschungsbereich “Ereignis Weimar-Jena—Kultur um 1800” installed in Jena University; the German Science foundation is thanked for major financial support. I thank Thomas Bach. Jan Frercks, Katja Regenspurger, Heiko Weber, und Gerhard Wiesenfeldt for providing data and for critical comments on an earlier version of this paper. Helen Lloyd Marshall helped with the English version of the manuscript. Universität Jena BIBLIOGRAPHY Bach, T. (2001a) “Dem Geist der Zeit eine neue Richtung geben” Die Naturphilosophie und die naturphilosophischen Professoren an der Universität Jena, In: Müller, G., Ries, K., Zische, P. (eds.) Die Universität Jena. Tradition und Fortschritt. Stuttgart: 155–174. Bach, T. (2001b) “Was ist das Thierreich anders als der anatomirte Mensch…?” Oken in Göttingen (1805–1807). In: Breidbach, Olaf; Fliedner, Hans-Joachim; Ries, Klaus (Hg.) Lorenz Oken (1779–1851). Ein politischer Naturphilosoph. Weimar: 73–91. Bach, T. (2001c) “Für wen das hier gesagte nicht gesagt ist, der wird es nicht für überflüssig halten.” Franz Joseph Schelvers Beitrag zur Naturphilosophie um 1800. In: Breidbach, Olaf; Ziche, Paul (Hg.) (2001) Naturwissenschaften um 1800. Wissenschaftskultur in Jena-Weimar. Weimar. Bach, T., Breidbach, O. (2001) “Die Lehre im Bereich der ,Naturwissenschaften’ an der Universität Jena zwischen 1788 und 1807”. NTM. 9: 152–176. Batsch, A. J. G. K. (1791) Versuch einer historischen Naturlehre oder einer allgemeinen und besonderen Geschichte der cörperlichen Grundstoffe. Für Naturfreunde entworfen. Zweiter physikalischer Theil. Halle. Bauer, Joachim, Gerd Müller (2001) “Zwischen Theologie und praktischen Wissenschaften: Der Aufklärer Joachim Georg Darjes.” In: Breidbach, O., Ziche, P. (eds) Naturwissenschaften um 1800. Wissenschaftskultur in Jena-Weimar: 142–154.

THE CULTURE OF SCIENCE AND EXPERIMENTS

213

Berger, J. (Hg.) (2001) Der Musenhof Anna Amalias. Geselligkeit, Mäzenatentum und Kunstliebhaberei im klassischen Weimar. Köln & Weimar. Bohley J. (2001) “Gemeinsame Interessen—wissenschaftliche Divergenzen? Die politischen Naturforscher Lorenz Oken und Christian Gottfried Nees von Esenbeck.” In Breidbach O., Fliedner H.-J., Ries K. (eds.), Lorenz Oken (1779–1851). Ein politischer Naturphilosoph. Weimar: 183–209. Breidbach, O. (1982) Das Organische in Hegels Denken. Würzburg. Breidbach, Olaf (1986) “Die Naturkonzeption Schellings in seiner frühen Naturphilosophie.” Phil. Naturalis 23: 82–95. Breidbach, Olaf (1988) “Schleidens Kritik an der spekulativen Naturphilosophie.” In Schleiden Matthias Jakob (Nachdruck) Schelling’s und Hegel’s Verhältnis zur Naturwissenschaft. Zum Verhältnis der physikalistischen Naturwissenschaft zur spekulativen Naturphilosophie. Weinheim:1–56. Breidbach, Olaf (1998a) Einleitung. In Schleiden, Matthias Jacob (Reprint) Grundzüge der Wissenschaftlichen Botanik. Hildesheim: V–XXIII. Breidbach O., (1998b) “Naturphilosophie und Medizin im 19. Jahrhundert.” In Pinkau K., Stahlberg C. (eds.), Deutsche Naturphilosophie und Technikverständnis. Stuttgart: S. Hirzel Verlag, 9–32. Breidbach, Olaf (1999a) “Zur Mechanik der Ontogenese.” In Wolff, C. F. (Reprint): Theoria Generationis. Frankfurt/a.M: I–XXXIV. Breidbach O. (1999b) “Prozessualität als systematische Kategorie in der Naturphilosophie Schellings.” In Leyte Coello A. (eds.), Una Mirada a la Filosofia de Schelling. Vigo: Servivio de Publicacións fe Universidade de Vigo: 147–160. Breidbach O. (1999c) “Zur Anwendung der Friesschen Philosophie in der Botanik Schleidens.” In Hogrebe W., Herrmann K. (eds.), Jakob Friedrich Fries, Philosoph, Naturwissenschaftler und Mathematiker. Frankfurt: Peter Lang, 221–242. Breidbach O. (2000) “Jenaer Naturphilosophen um 1800,” Sudhoffs Archiv für Geschichte der Medizin und der Naturwissenschaften, 84:19–49. Breidbach, O. (2001a) “Transformation statt Reihung—Naturdetail und Naturganzes in Goethes Metamorphosenlehre.” In: Breidbach, O., Ziche, P. (eds) Naturwissenschaften um 1800. Wissenschaftskultur in Jena-Weimar. Weimar: 46–65. Breidbach O. (2001b) “Oken in der Wissenschaftsgeschichte des 19. Jahrhunderts.” In Breidbach O., Fliedner H.-J., Ries K. (eds.), Lorenz Oken (1779–1851). Ein politischer Naturphilosoph. Weimar: 15–32. Breidbach, O. (2003) Schelling und die Erfahrungswissenschaften. Submitted to Sudhoffs Archiv. Breidbach, O., Frigo, G. F. (2003) Zur Rezeption der Naturforschung des mitteldeutschen Raumes an der Universität Padua zwischen 1770 und 1820: Sudhoffs Archiv für Geschichte der Medizin und der Naturwissenschaften: In press Breidbach, O., Frigo, G. F., Piovan, F. (2003) Naturforschung im Lehrangebot der Universität Padua zwischen 1770 und 1815. NTM: In press. Breidbach, O., Ghiselin, M. (2003) Lorenz Oken’s Naturphilosophie in Jena, Paris and London. HPLS: In press. Breidbach O., Ziche P. (2001) ‘Naturwissen und Naturwissenschaften. Zur Wissenschaftskultur in Weimar/Jena’. In Breidbach, O., Ziche, P. (eds) Naturwissenschaften um 1800. Wissenschaftskultur in Jena-Weimar. Weimar:7–24. Clark, W. (1997) ‘German Physics textbooks in the Goethezeit’ Part 1, 2. The British Journal for the History of Science 35: 219–239, 295–363. Charpa, Ulrich (1989) Zur Einführung. In: Matthias Jakob Schleiden, Wissenschaftsphilosophische Schriften. Köln: 9–44. Chemnitius, F. (1929) Die Chemie in Jena von Rolfinck bis Knorr. Jena. Cunningham A., Jardine N. (eds.) (1990) Romanticism and the Sciences, Cambridge. Dieckmann A. (1992) Klassifikation-System-‘scala naturae’ das Ordnen der Objekte in Naturwissenschaft und Pharmazie zwischen 1700 und 1850, Stuttgart. Döbling, H. (1928) Die Chemie in Jena zur Goethezeit. Beiheft 13, Zeitschrift des Vereins für Thüringische Geschichte und Altertumskunde. Jena. Engelhardt, Dietrich von. (1992) “Müller und Hegel. Zum Verhältnis von Naturwissenschaft und Naturphilosphie im deutschen Idealismus.” In Hagner, Michael; Wahrig-Schmidt, Bettina (Hg.) Johannes Müller und die Philosophie. Berlin 85–103. Engelhardt D. von. (1997) “Naturforschung im Zeitalter der Romantik.” In Zimmerli W. C., Stein K., Gerten M. (eds), Fessellos durch die Systeme. Frühromantisches Naturdenken im Umfeld von Arnim, Ritter und Schelling. Stuttgart: 19–48. Erxleben, J.C. P. (1796) Anfangsgründe der Naturlehre. Frankfurt und Leipzig. Fischer, J. K (1797) Anfangsgründe der Physik in ihrem mathematischen und chemischen Theile nach den neuesten Endeckungen. Jena.

214

O. BREIDBACH

Fischer, Johann Carl (1798) Physikalisches Wörterbuch oder Erklärung der vornehmsten zur Physik gehöriger Begriffe. Erster Theil. Göttingen. Fischer, J. C. (1801–1808) Geschichte der Physik seit dem Wiederaufbau der Wissenschaften bis Ende des 18. Jahrhunderts Vol. 1–8: Göttingen. Frercks, J. (2003) “Disziplinbildung und Vorlesungsalltag. Funktionen von Lehrbüchern der Physik um 1800 mit einem Fokus auf die Universität Jena.” Berichte zur Wissenschaftsgeschichte. In press. Fries J. F. (1822) Die mathematische Naturphilosophie nach philosophischer Methode bearbeitet. Ein Versuch, Heidelberg. Genschorek, W. (1977) Christoph Wilhelm Hufeland. Der Arzt der das Leben verlängern half. Leipzig. Ghiselin M. T. (2000) ‘The founders of morphology as alchemists.’ In Ghiselin M.T., Leviton A.E. (eds.), Cultures and Institutions of Natural History: Essays in the History and Philosophy of Science. San Francisco: California Academy of Sciences, 39–49. Giese E., Hagen B. von (1958) Geschichte der Medizinischen Fakultät der Friedrich-Schiller-Universität Jena, Jena. Glasmachern, Thomas (ed.) (1989) Fries—Apelt—Schleiden. Verzeichnis der Primär- und Sekundärliteratur 1798–1988. Köln. Hackman, W. (1978) Electricity from Glass. The History of the Frictional Electrical Machine 1600– 1850. Alphen aan den Rijn. Hammerstein, N. (2001) ‘Die deutsche Universitätslandschaft im ausgehenden 18. Jahrhundert.’ In Müller, G., Ries, K., Ziche, P. (eds) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 13–26. Henrich, D. (1991) Konstellationen, Probleme und Debatten am Ursprung der idealistischen Philosophie (1789–1795) Stuttgart. Herz, H. (1989) Von Schillers Berufung bis Fichtes Entlassung. Vorlesungen an der philosophischen Fakultät der Universität Jena 1789–1799. Jena. Hinske, N., Lange, E, Schröpfer, H. (eds.)(1994) “Das Kantsche Evangelium.” Der Frühkantianismus an der Universität Jena von 1785–1800. Stuttgart. Humboldt, Alexander von (1797) Versuche über die gereizte Muskel- und Nervenfaser: nebst Vermuthungen über den chemischen Process des Lebens in der Thier- und Pflanzenwelt. Posen and Berlin. Jahn, I. (1963) Geschichte der Botanik in Jena von der Gründung bis zur Berufung Pringsheims (1558–1864) Diss. (typescript) Jena. Kant, I. (1786) Metaphysischen Anfangsgründe der Naturwissenschaft. Klengel, B. (2001) ‘Ultraviolette Strahlen und Ladungssäule. Zur Rezeption von Entdeckungen Johann Wilhelm Ritters in Frankreich.’ In Breidbach, O., Ziche, P. (eds.) Naturwissenschaften um 1800. Wissenschaftskultur in Jena-Weimar. Weimar: 248–258. Knight, D. (1981) Ordering the World. A History of Classifying Man. London. Knight, D. (1984) The Age of Science. The Scientific World-View of the Ninetheenth Century. Oxford Knight D. (1998) Science in the Romantic Era. Aldershot: Ashgate. Krause, Karl Christian Friedrich (1804) Entwurf eines Systems der Philosophie. Jena and Leipzig. Krehr-Hartmann, B. (1999) “Zur Struktur der Sammlung der” Societät für die gesammte Mineralogie zu Jena um 1800”. European Journal of Mineralogy. Supplement 1 Kuhn D. (1988) Typus und Metamorphose. Goethe Studien, Marbach: Deutsche Schillergesellschaft. Lepenies W. (1976) Das Ende der Naturgeschichte: Wandel kultureller Selbstverständlichkeiten in den Wissenschaften des 18. und 19. Jahrhunderts, München. Lenoir, T (1978) “Generational factors in the origin of Romantische Naturphilosophie” Journal for the History of Biology 11: 57–100. Lenoir, T. (1982) The Strategy of Life: Teleology and Mechanics in Nineteenth-Century German Biology, Utrecht: D. Reidel Lichtenberg; G. C. (1778) erste Abhandlungen allgemeine Experimente enthaltend über eine neue Methode, die Natur und die Bewegung der elektrischen Materie zu erforschen. Vorgetragen in der öffentlichen Sitzung der Königlichen Gesellschaft der Wissenschaften am 21. Februar 1778. Göttingen. Lind, G. (1992) Physik im Lehrbuch 1700–1850. Berlin. Linné, A. (In prep.) The natural families of A. J. G. K. Batsch—the silent takeover of natural history by systematics. Manger; K. “Aufklärung in der Literaturwissenschaft. Ereignis Weimar_Jena”. In Enskat, R. (ed.) Wissenschaft und Aufklärung, Opladen 97–118. Mayer, J. T. (1801) Anfangsgründe der Naturlehre zum Behuf der Vorlesungen über die ExperimentalPhysik. Göttingen.

THE CULTURE OF SCIENCE AND EXPERIMENTS

215

Martens, F. H. (1803) Vollständige Anweisung zur therapeutischen Anwendung des Galvanismus. Weißenfels & Leipzig. Moiso F. (1994) “Magnetismus, Elektrizität, Galvanismus”. In Baumgartner H., Jacobs W. G., Krings H. (eds.), Historisch-kritische Ausgabe im Auftrage des Schelling Kommission der Bayerischen Akademie der Wissenschaften. Stuttgart: Fromann und Holzbach: Ergänzungsband zu 1.5. Wissenschaftshistorischer Bericht zu Schellings naturphilosophischen Schriften 1797–1800., 165–372. Müller, G. (2001) ‘Perioden Goethescher Universitätspolitik’. In Müller, G., Ries, K., Ziche, P. (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 135–154. Müller, G. (2001a) Die extraordinäre Universität. Jenas Modernisierungsweg. In Müller, G., Ries, K., Ziche, P. (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 191–196. Müller, K. D. (1992) Franz Joseph Schelver 1778–1832. Stuttgart. Müller-Wille S. (2000), Botanik und weltweiter Handel. Zur Begrundung eines Natürlichen Systems der Pflanzen durch Carl von Linné (1707–1778), Berlin: Verlag für Wissenschaft und Bildung. Neuser W. (1995) Natur und Begriff, Stuttgart. Neuser, W. (1997) ‘Die Methoden der Naturwissenschaften im Spiegel der frühen Naturphilosophie Schellings’. In Zimmerli W. C., Stein K., Gerten M. (eds.), Fessellos durch die Systeme. Frühromantisches Naturdenken im Umfeld von Arnim, Ritter und Schelling. Stuttgart: 331–340. Pera, M. (1992) The Ambiguous Frog. The Galvani-Volta Controversy on Animal Electricity. Princeton, NJ. Poggi S. (2000) Il Genio e l’Unità della Natura. La Scienza della Germania Romantica (1790–1830), Bologna. Polianski, I. (2001) “Der außerordentliche Garten. Zur Geschichte des Herzoglichen Botanischen Gartens zu Jena.” In Müller, G., Ries, K., Ziche, P. (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 205–220. Rasche, U. (2001) “Umbrüche—Zur Frequenz der Universität Jena im ausgehenden 18. und frühen 19. Jahrhundert”. In Müller, G., Ries, K., Ziche, P., (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart 79–134 Richards, Robert J. (1992) The Meaning of Evolution. The Morphological Construction and Ideological Reconstruction of Darwin’s Theory. Cambridge. Richter, K. (ed.) (1988) Der Physiker des Romantikerkreises Johann Wilhelm Ritter in seinen Briefen an den Verleger Carl Friedrich Ernst Frommann. Weimar. Richter, K. (1997) “Zur Methodik des naturwissenschaftlichen Forschens bei Johann Wilhelm Ritter”. In Zimmerli W.C., Stein K., Gerten M. (eds), Fessellos durch die Systeme. Frühromantisches Naturdenken im Umfeld von Arnim, Ritter und Schelling. Stuttgart: Frommann-Holzboog: 317–330. Roe, S. A. (ed.) (1981) The Natural Philosophy of Albrecht von Haller. New York Rupke N. A. (1994) Richard Owen: Victorian Naturalist, New Haven, CT: Yale University Press. Schad, J. B. (1804) System der Natur- und Transscendentalphilosophie in Verbindung dargestellt. Zweiter Theil. Landshut. Schäffner, K. (2001) “Lorenz Oken in Zürich (1832–1851)”. In Breidbach, O., Fliedner H.-J,m Ries, K. (eds.) Lorenz Oken 1779–1851. Ein politischer Naturphilosoph. Weimar: 120–145. Schelling, F. W. J. (1799) “Einleitung zu dem Entwurf eines Systems der Naturphilosophie oder über den Begriff der spekulativen Physik und die innere Organisation eines Systems dieser Wissenschaft”. In Schröter M. (eds.), Schellings Werke (1927), vol. 2. München: Beck, 269–327. Schleiden, Matthias Jacob (1842) Grundzüge der wissenschaftlichen Botanik nebst einer methodologischen Einleitung als Anleitung für das Studium der Pflanze. I. Theil. Leipzig. Schleiden, Matthias Jakob (1844) Schelling’s und Hegel’s Verhältnis zur Naturwissenschaft. Leipzig. Schmid I. (1979) Die Oberaufsicht über die naturwissenschaftlichen Institute bei der Universität Jena unter Goethes Oberaufsicht. Ein Beitrag zur Geschichte der Oberaufsicht über die unmittelbaren Anstalten für Wissenschaft und Kunst in Sachsen-Weimar-Eisenach. Thesis. Berlin. Schmidt, G. C. (1773) Beschreibung einer Elektrisir-Maschine und deren Gebrauch. Jena. Sichau, C. (2000) “Die Replikationsmethode: Zur Rekonstruktion historischer Experimente”, In Heering, P., Rieß, F., Sichau, C. (eds.) Im Labor der Physikgeschichte. Oldenburg: BIS, 2000. Steffens, H. (1818) Ueber die elektrischen Fische. Frankfurt a. M. Steiger, G. (1958) “Die Universität Jena im Zeitalter der Befreiungskriege, der Restauration und des Kampfes um die Deutsche Einheit (1806–1871). Die allgemeine Entwicklung der Universität zwischen der Schlacht von Jena und den Beschlüssen von Karlsbad (1806–1819/21).” In Steinmetz, M. (ed.) Geschichte der Universität Jena 1548/58–1958. Jena: 319–359. Steiger, G. (1986) Goethe, die Universität Jena und die Naturwissenschaften. Jena Stevens P. E. (1994) The Development of Biological Systematics. Antoine-Laurent de Jussieu, Nature and the Natural System, New York: Columbia University Press.

216

O. BREIDBACH

Stichweh, R. (1984) Zur Entstehung des modernen Systems wissenschaftlicher Disziplinen. Physik in Deutschland 1740–1890. Frankfurt a.M. Stichweh, R. (1994) Wissenschaft, Universität, Professionen. Soziologische Analysen. Frankfurt a.M. Sudhoff, K. (1922) Hundert Jahre Deutsche Naturforscher-Versammlungen. Gedächtnisschrift zur Jahrhundert-Tagung der Gesellschaft Deutscher Naturforscher und Ärzte, No place given. Voigt, J. H. (1793) Versuch einer neuen Theorie des Feuers, der Verbrennung, der künstlichen Luftarten, des Athmens, der Gährung, der Electricität, der Meteoren, des Lichts und des Magnetismus. Aus Analogien hergeleitet und durch Versuche bestätigt. Jena. Voigt, F. S. (1808) System der Botanik. Jena Walther, G. (2001) “Das Ideal: Göttingen. Ruf, Realität und Kritiker der Georgia Augusta um 1800.” In Müller, G., Ries, K., Ziche, P. (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 33–46. Walther, H. G. (2001) “Die Universität Jena um 1800.” In Müller, G., Ries, K., Ziche, P. (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 27–32. Weber, H. (2002) Die Elektrisiermaschine im 18. Jahrhundert. Ernst Haeckel Haus Studien- Monographien zur Geschichte der Biowissenschaften und Medizin Bd. 7. Berlin. Weber, H., Breidbach, O., Heide, K., Grimm, H. (2003) Glas und Elektrizität. Zur Glastechnologie in der Konstruktion der Weimarer Elektrisiermaschine von 1773. Glass, Science and Technology. Glastechnische Berichte: In press. Weber, H., Breidbach, O. (in prep.) Überlieferung und Bestand. Zur Geschichte der Elektrisiermaschinen bis 1800. Wiesenfeldt, G. (2003) “Selbstrezeption und Fremdrezeption: Die Selbstexperimente Johann Wilhelm Ritters.” In Bach, T., Breidbach, O., Wiesenfeldt, G. (eds.) Rezeption und Identitäten: Methodische Zugänge zur Wissenschaftkultur um 1800. In prep. Winckler, J. H. (1745) Die Eigenschaften der electrischen Materie und des electrischen Feuers aus verschiedenen neuen Versuchen erkläret, und nebst etlichen neuen Maschinen zum Electrisieren beschrieben. Leipzig. Wyder, M. (1998) Goethes Naturmodell. Die Scala naturae und ihre Transformationen. Köln and Weimar. Zaunstöck, H. (2001) “Untersuchungen zur Struktur Naturforschender gesellschaften im 18. Jahrhundert. Die Sozietäten in Halle, Leipzig und Jena”. In Breidbach, O., Ziche, P. (eds.) Naturwissenschaften um 1800. Wissenschaftskultur in Jena-Weimar. Weimar: 155–178. Ziche, Paul (1998) “Von der Naturgeschichte zur Naturwissenschaft. Die Naturwissenschaften als eigenes Fachgebiet an der Universität Jena.” Berichte zur Wissenschaftsgeschichte 21: 251–263. Ziche, P. (2001) “Die Grenzen der Universität. Naturforschende, physikalische und mechanische Aktivitäten in Jena.” In: Müller, G., Ries, K., Ziche, P. (eds.) Die Universität Jena. Tradition und Innovation um 1800. Stuttgart: 221–237. Ziche, P., Büch, G., Kenklies, K., Neuoer, H., Breidbach, O., 2000, “Eine naturwissenschaftliche Forschungsbibliothek des 18. Jahrhunderts: die Bibliothek de r ‘Naturforschenden Gesellschaft’ zu Jena,” Berichte zur Wissenschaftsgeschichte, 23: 433–447.

THE ROMANTIC EXPERIMENT AS FRAGMENT ROBERT MICHAEL BRAIN

Viele der ersten Stifter der modernen Physik müssen gar nicht als Philosophen, sondern als Künstler betrachtet werden. —Friedrich Schlegel, Athenäums-Fragmente

1. INTRODUCTION Historians of science often refer loosely to Naturphilosophie as “romantic”. Strictly speaking, however, there was a certain intellectual gap between the systematic philosophical thinkers and their romantic counterparts whose work they so deeply conditioned and shaped. Even in cozy Jena the philosophers stood at a certain intellectual and social distance from the romantic circles of the Schlegels, Novalis, Tieck, and others: Schelling did not write for the Athenäum, nor did Fichte and Hegel.1 Nor did the philosophers write much of significance in the literary genres usually thought to be most characteristic of romanticism: the novel, poem, and fragment. Nevertheless, philosophy—the different idealistic philosophies of Kant, Fichte, and Schelling—served as a condition for the possibility of early romanticism. But the romantics parted company with Schelling’s philosophical idealism in their conviction that a purely theoretical completion of the System is impossible, that its infinite progressive process is asymptotic to its infinitely distant goal. Instead, the romantics pursued the philosophical quest in the empirical world, seeking the Absolute in the reflexive objects of history, nature, and especially, the particular and finite artwork. As Walter Benjamin wrote in his study of art criticism in German romanticism, “the absolute… in the period of the Athenäeum was in fact the figure [Gestalt] of art. But …[one] did not seek this Absolute systematically: quite the inverse, one sought to grasp the System ‘absolutely’.”2 This paper contends that Ørsted, and to some degree other key natural scientists associated with the Jena early romantics—Johann Wilhelm Ritter, Novalis

1

2

Throughout this article I will use the word “romantic” to refer to the so-called “early romantics” in Jena, as as a few of their direct intellectual descendents, such as Lorenz Oken and Robert Schumann. Walter Benjamin, Der Begriff der Kunstkritik in der deutschen Romantik (Frankfurt am Main: Suhrkamp, 1973), p. 40.

217 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 217–233. © 2007 Springer.

218

R. M. BRAIN

(Friedrich von Hardenberg), Alexander von Humboldt—positioned the task of experiment vis-à-vis Naturphilosophie as a direct parallel to the Romantic artwork. With this claim I wish to address the abiding conundrum of the relation between theory and experiment among the alleged romantic scientists. Although the question has arisen with regard to the work of several experimentalists the debate has been most heated in the case of Ørsted, whose enduring scientific achievements and seemingly mixed messages on the philosophy of experiment have helped raise the stakes higher than for other figures. I take for granted the claim first made by Andrew Wilson, and extended by Michael Friedman (in this volume), that the involvement of mature Kantian and Schellingian dynamical theory with experimental practice arose from a very specific and fortuitous encounter between the philosophy of nature and new experimental discoveries in electrochemistry.3 But this paper contends that there was a third term in the mix, namely the interventions of the so-called “early Romantics” of Jena, particularly the circles which included the Schlegels, Novalis, and the authors who contributed to the Athenäum. Besides the specific and rigorously philosophical grounds which tied chemical experiment to speculative reason, aspects of the particular approach to experiment were informed by romantic aesthetic theory and criticism. “Don’t forget that we are artists!,” Ritter wrote in a letter to Ørsted, adding that “I hardly need to define art for you… 4 Ritter claimed the artists’ mantle systematically in his lecture “Physics as an Art,” where he defined art as humans’ expressive use of their intellectual faculties or life force for the purpose of creating self-awareness.5 In this lecture, Ritter arrayed the arts in hierarchical order, with physics at the top, just above music. Ørsted similarly described his work as the “physics of beauty” and frequently related scientific investigation to aesthetic experience. One might think that these terms were little more than schöngeisterei, attempts to grab a piece of the action in an age when artists were king. But this would be wrong. For Ritter and Ørsted such terms indicated the fundamental way in which their notion of experiment followed the philosophical logic of the work of art in romanticism. Many commentators of the period—Novalis foremost among them—regarded observation of natural objects and immersion in the artwork as the same reflexive process applied to different objects.6 Art criticism was a kind of experiment on the artwork, an awakening of the power of reflexion through the work, which brought self-consciousness and self- knowledge.7 Ørsted wrote similarly that “Experimentation… is the true art of the physicist, and if he has thus, with open eyes, really seen 3

4

5 6 7

Andrew Wilson, “Introduction,” Jelved, Jackson, and Knudsen (eds), Selected Scientific Writings of Hans-Christian Ørsted (Princeton, NJ: Princeton University Press, 1998); and Michael Friedman, “Kant—Naturphilosophie—Electromagnetism,” (in this volume). Ritter to Ørsted, 16 and 17 August 1805, in M. C. Harding (ed.), Correspondance de H. C. Örsted avec divers savants, 2 vols. (Copenhagen: H. Aschehoug & Co., 1920), p. 114. On Ritter and Ørsted’s artistic aspirations and the Jena context see also Dan Ch. Christensen, “Physics as a Branch of Art: The Romantics in Jena,” in Mogens Bencard (ed.) Intersections. Art and Science in the Golden Age (Copenhagen: Gyldendal, 2002). Johann Wilhelm Ritter, Die Physik als Kunst (Muenchen: J. Lindauer, 1806). Benjamin, Der Begriff der Kunstkritik … pp. 48–56. Novalis: “Was zugleich Gedanke und Beobachtung ist, ist ein kritischer… Keim.” See Benjamin, Der Begriff der Kunstkritik…, p. 60.

THE ROMANTIC EXPERIMENT AS FRAGMENT

219

part of nature reconstruct itself, he can, from this point, survey or, at least, sense the coherence of all nature.” Historians of science have spilled a river of ink over whether Ørsted and his friends drew meaningful philosophical guidance from Schelling, in particular, in their quest to discover the hidden relations between natural forces.8 Ørsted frequently said that he had, of course, and most historians now take him more or less at his word.9 The difficulty arises because Ørsted’s tributes to Schelling often also draw limits to the role of philosophical speculation, asserting that it can only be completed through empirical investigation. In a typical passage, Ørsted wrote that “it is well-known that Schelling, through speculation, has produced an attempt which, as such, is of incalculable value, but the combined efforts of a great number of blessed geniuses are probably required for the accomplishment of this task.”10 Historians of science have understandably stumbled on these remarks since on Schelling’s aprioristic terms, knowledge of natural objects should be derivable from reason alone. In what follows, I will sketch the terms of the romantic answer to the shortcomings of speculative reason, arguing that Ørsted’s notion of experiment was similarly conceived. Like the artist-philosophers who grasped the Absolute in the figure of art, the experimentalists found it in the figure of the artful experiment.

2. ØRSTED AND THE EARLY ROMANTICS Ørsted’s literary and aesthetic interests began early and never ceased.11 He published poetry, wrote essays on aesthetics, edited a monthly literary magazine for nine years, reviewed theater and opera, and wrote literary and philosophical dialogues over a span of four decades. Ørsted’s contact with the work of the Jena authors 8

9

10

11

The highlights of this literature begin with Robert Stauffer’s “Speculation and Experiment in the Background of Ørsted’s Discovery of Electromagnetism,” Isis 48 (1957), pp. 33–50, whose arguments were made central to 19th-century physics by Thomas Kuhn, “Energy Conservation as an Example of Simultaneous Discovery,” in M. Clagett (ed.), Critical Problems in the History of Science (Madison, WI: University of Wisconsin Press, 1959). It was further extended by L. Pearce Williams in The Origins of Field Theory (New York: Random House, 1966) and “Kant, Naturphilosophie and Scientific Method,” in R. Giere and R. Westfall (eds.), Foundations of Scientific Method in the Nineteenth Century (Bloomington, IN: Indiana University Press, 1973). Barry Gower challenged these readings in “Speculation in Physics: The History and Practice of Naturphilosophie,” Studies in History and Philosophy of Science 3 (1973), pp. 301–356; as did Timothy Shanahan, “Kant, Naturphilosophie, and Ørsted’s Discovery of Electromagnetism: A Reassessment,” Studies in History and Philosophy of Science 20 (1989), pp. 287–305. Kenneth Caneva gives a careful and nuanced affirmation in “Physics and Naturphilosophie: A Reconaissance,” History of Science 35 (1997), pp. 35–106; Michael Friedman gives a philosophically precise account in “Kant—Naturphilosophie—Electromagnetism,” in this volume. For recent dissenting views see Dan Ch. Christiansen, “The Ørsted–Ritter Partnership and the Birth of Romantic Natural Philosophy,” Annals of Science 52 (1995), pp. 153–185; and Maria Jean Trumpler, “Questioning Nature: Experimental investigations of animal electricity in Germany, 1791–1810,” Ph.D. dissertation Yale University, May 1992. Hans-Christian Ørsted, “New Investigations into the Question: What is Chemistry? (1805),” in Jelved, Jackson, Knudsen ( transl. and eds.), Selected Scientific Writings of Hans Christian Ørsted …, p. 199. Wilson’s, “The Unity of Physics and Poetry: H. C. Ørsted and the Aesthetics of Force,” Journal of the History of Ideas…,” provides a fine account of several key phases of Ørsted’s literary and poetic activities.

220

R. M. BRAIN

began during his student years, and it later developed through personal ties and friendship. In an 1807 letter to the Danish poet Adam Oehlenschläger, Ørsted remembered that he was already “a friend of the new philosophy and poetry” when the two met in 1797, while Oehlenschläger at the time dismissed the Schlegel brothers as “impudent youths.”12 When the Schlegel brothers came out with the journal Athenäum from 1798 to 1800, the principal organ of early romanticism, Ørsted devoured it and prepared to travel to Germany to meet Friedrich Schlegel and discuss his aesthetic ideas.13 During his European tour in 1801–1802, Ørsted attended Friedrich Schlegel’s lectures on aesthetics in Berlin during the Fall and Winter semester, and also developed cordial personal relations with him, which included dining together frequently and joining him in visits to the Berlin salon of Henriette Herz. In a diary entry from early 1802, Ørsted noted that he expected “to profit much from [Schlegel’s] friendship.” And indeed he did. Besides the edification he derived from personal conversation with the doyen of German romanticism, Schlegel introducing Ørsted to the broader Jena circles, including Novalis (Friedrich von Hardenberg) and Ritter. When Ørsted journeyed to Paris several months later, he met up again with Schlegel, who had decamped to the French capital. In Paris Ørsted joined a group of Germans who met on Sundays at Schlegel’s home to converse “about philosophical, physical, or aesthetic subjects.”14 Ørsted passionately shared the literary and philosophical preoccupations of the Schlegel circle, including the common concern to integrate physics with aesthetic pursuits. After returning to Denmark, Ørsted fiercely defended Schlegel and Novalis against their Danish detractors.15 These biographical details should amend the picture that predominates in accounts of Ørsted’s first visit to Germany. Historians of science typically frame the terms of the philosophical discussion by noting that Ørsted briefly attended Schelling’s lectures in Jena, and began his intense scientific association with Ritter.16 But the novelty of Ørsted’s German sojourn lay not in his exposure to speculative 12

13

14 15 16

Ørsted, quoted in Andrew Wilson, “The Unity of Physics and Poetry …”, quotation taken from a letter written by Ørsted to the Danish poet Adam Oehlenschlaeger some years later, on November 1, 1807, published in Breve fra og til Hans Christian Ørsted, edited by Mathilde Ørsted, 2 vols., (Kjobenhavn: Th. Linds Forlag, 1870), vol. 1, p. 226. For Ørsted’s familiarity with the Atheneum see Breve…, vol.1, p. 56 and Anders Sandoe Ørsted, Af Mit Livs og Min Tids Historie, forkortede Udgave (København: Arne-Frost-Hansens Forlag, 1951), pp. 40, 138–139. Wilson also cites Ørsted praising Novalis’s “great and free overview” of nature and “well-ordered knowledge of the most important parts of all the sciences,” and declaring that the romantic’s poetic and scientific works had given him “so many richly enjoyable hours.” Wilson, ibid, p. 8. I thank Andrew Wilson for providing me with the manuscript of his splendid paper, which contains further Danish archival references to Ørsted’s literary relations with the Jena circle. The Athenaeum, which appeared twice a year between 1798 and 1800, was the theoretical organ of the early Romantic movement in Germany. It was edited by Friedrich and August Wilhelm Schlegel, and included contributions by Novalis, Friedrich Schleiermacher, and the editors. In 1803 Friedrich Schlegel founded the journal Europa as the successor to the Athenaeum, to which Ørsted contributed. In the first issue of Europa, Schlegel described the transition as follows: “In the early issues [of the Athenaeum], critique and universality are the primary goal; in the later parts, the spirit of ‘mysticism’ is essential. One shouldn’t shrink from using this word.” Schlegel, quoted in Benjamin, Der Begriff der Kunstkritik…, pp. 90–91. Ørsted, Breve, vol. 1, 37, quoted in Wilson, “The Unity of Physics and Poetry.…” Wilson, “The Unity of Physics and Poetry…” On the Ritter-Ørsted collaboration see Dan Ch. Christensen, “The Ørsted-Ritter Partnership and the Birth of Romantic Natural Philosophy,” Annals of Science 52 (1995), pp. 153–185.

THE ROMANTIC EXPERIMENT AS FRAGMENT

221

idealism—he had already staked out a strongly speculative reading of the late Kant in his two dissertations, perhaps with some encouragement from his 1797 reading of Schelling’s Ideas—but in his experience with galvanic phenomena and the experiments with it that he undertook after meeting Ritter.17 These took place not simply within a Schellingian framework, but within a larger discussion about the meaning of Schelling’s philosophy for both art and the empirical philosophy of nature that was being carried out by Friedrich Schlegel and his circle, which included both Novalis and Rittter. They were all avid, yet partially dissatisfied readers of Schelling, of course, as well as of Kant and Fichte. But as I have suggested, and will argue further below, they read the Idealist philosophy in a distinct way, and derived from it a specific agenda. With this in mind, one might interpret Ørsted’s collaboration with Ritter differently than historians of science have typically done. Rather than asking simply whether their galvanic experiments were or were not attempts to apply Schelling’s Naturphilosophie to real experiments, I propose that we ask whether they were attempts to conceptualize experiment anew in light of the philosophical and aesthetic concerns of the Jena circle.18

3. ROMANTIC VERSUS NEWTONIAN EXPERIMENT The romantic experimentalists signaled their break with the dominant philosophy of experimental sciences by challenging the hallowed experimentum crucis.19 Francis Bacon had called crucial experiments “instances of the fingerposts,” which in his words, “afford very great light, and are of high authority, the course of interpretation sometimes ending in them and being completed.”20 But it was Newton, of course, who formulated the canonical version of the experimentum crucis, in his famous prism trials showing that the sun’s light is composed of rays which are differentially refrangible.21 Newton argued that amongst the many trials an experimentalist might conduct, it was possible to pick out those that were 17

18

19

20 21

For Ørsted’s reading of Kant, see Friedman, “Kant–Naturphilosophie– Electromagnetism,” (in this volume), [p.] and especially fn. 40. As a biographical matter, it is worth noting that Ritter’s letters to Ørsted (we only have one side of the correspondence) frequently included greetings from Schlegel, or, when Ørsted was with Schlegel in Paris, offer greetings to be extended through Ørsted. The overall impression is that these men understood themselves as a close-knit circle of friends. See Ritter’s letters to Ørsted in M.C. Harding (ed.), Correspondence de H.C. Örsted …, pp. 8, 31, 40, 43, 45, 51, 57. For recent reappraisals of crucial experiments, see Ian Hacking, Representing and Intervening (Cambridge: Cambridge University Press, 1983), pp. 246–261. Bacon, quoted in Hacking, Representing …, p. 250. Isaac Newton, Opticks, or a Treatise of the Reflections, Refractions, Inflections & Colours of Light (New York: Dover, 1952 [based on the fourth edition published by William Innys of London in 1730], pp. 32–33. On Newton’s crucial experiment and its place in its optics see Z. Bechler, “Newton’s 1672 Optical Controversies. A study in the grammar of scientific dissent,” in Y. Elkana (ed.), The Interaction between Science and Philosophy (Atlantic Highlands: Humanities Press, 1974), pp. 115– 142; Bechler, “A less agreeable matter.” The disagreeable case of Newton and achromatic refraction,” British Journal for the History of Science 8 (1975), 101–126; Geoffrey Cantor, “The Rhetoric of Experiment,” The Uses of Experiment. Studies in the natural sciences (Cambridge: Cambridge University Press, 1989), pp. 159–180; Peter Dear, Totius in verba: Rhetoric and authority in the early Royal Society,” Isis 76 (1985), pp. 145–161; Simon Schaffer, “Glass works,” in David Gooding, Trevor Pinch, and Simon Schaffer (eds.), The Uses of Experiment…, pp. 67–104.

222

R. M. BRAIN

“crucial” and would prove the truth or falsity of a hypothesis. For Newton and his followers, moreover, the crucial experiment illustrated the proper boundary between experiments, “concluding directly and without any suspicion of doubt”, and hypotheses, “conjectured by barely inferring ‘tis thus because not otherwise or because it satisfies all phaenomena.”22 Proper experimentation, in other words, delineated all imaginable hypotheses and demonstrated the incontestability of one, thereby bringing closure to the question. Thus, an experiment attains individuation, acquires a unique identity and unarguable meaning. Experimentalists of the early romantic generation regarded the Newtonian doctrine with suspicion. No doubt Goethe’s famous challenges to Newtonian experiment shaped their attitudes.23 In his various writings on optics and color theory, culminating in the Farbenlehre, Goethe contended that the experimentum crucis was far from self-evident. The unarguable meaning which attached to Newton’s experiments, he contended, was largely a rhetorical accomplishment, achieved through struggle, and shored up by the social and institutional power of Newtonian alliances.24 But it was not just the politics of Newtonian science which galled Goethe. It was the way in which Newton privileged the theory of refraction over all other views of light and color, and thus upheld a limited or even artificial relation to nature. Ørsted, at least in the published writings I’m aware of, spared the political polemics, but he endorsed the epistemological critique of Newton’s philosophy of experiment. “It is inherent in the infinitude of Nature that no observer can discover all that is implied by an experiment,” he wrote, continuing: To understand an experiment completely would be equivalent to having found the key to all of Nature. Therefore, the perspicacious discoverer of acoustic figures cannot be blamed if he has not observed all that is really implied by his experiments. Did not Newton himself in his masterly enquiries into prismatic colours overlook several quite important phenomena, which also escaped the attention of his successors for a whole century until Hershel’s (sic) and Ritter’s experiments enlightened us?25

Ørsted, like other romantic experimentalists, believed Nature to be infinite and “an activity which knows no rest.”26 The possible relevant phenomena are so varied

22

23

24

25

26

Isaac Newton, Correspondence, 7 vol., (eds.) H.W. Turnbull, J. F. Scott, A. R. Hall and L. Tilling (Cambridge: Cambridge University Press, 1959–1977), I, pp. 96–97. On relations between Goethe and the Jena romantics see Frederick Beiser, German Idealism: The Struggle Against Subjectivism (Cambridge, MA: Harvard University Press, 2002) ; Robert J. Richards, The Romantic Conception of Life: Science and Philosophy in the Age of Goethe (Chicago, IL: University of Chicago Press, 2002). J. W. Goethe, Farbenlehre: Theoretischer Schriften (Stuttgart: Kohlhammer, 1953). On Goethe’s polemics, see Myles W. Jackson, “A Spectrum of Belief: Goethe’s ‘Republic’ versus Newtonian ‘Despotism’,” Social Studies of Science 24 (London, Thousand Oaks, and New Delhi: SAGE, 1994), pp. 673–701; Frederick Burwick, The Damnation of Newton: Goethe’s Color Theory and Romantic Reception (Berlin and New York: Walter de Gruyter, 1986); Dennis Sepper, Goethe Contra Newton: Polemics and the Project for a New Science of Color (Cambridge, New York, New Rochelle, Melbourne and Sydney: Cambridge University Press, 1987); Ruppert Matthaei, Goethe’s Farbenlehre (Ravensburg: Otto Maier Verlag, 1971); “Experiment,” Goethe Wörterbuch (ed.) Akademie der Wissenschaften der (vormaligen) DDR/Akademie der Wissenschaften in Goettingen/Heidelberger Akademie der Wissenschaften, Bd. 3 (Stuttgart, 1993), pp. 499–502. Ørsted, “Experiments on Acoustic Figures,” in Jelved, Jackson, and Knudsen (eds.), Selected Scientific Works of Hans Christian Ørsted…, p. 266. Ørsted, “First Introduction to General Physics (1811),” in Jelved, Jackson, and Knudsen, Selected Scientific Works of Hans Christian Ørsted …, p. 284.

THE ROMANTIC EXPERIMENT AS FRAGMENT

223

and so numerous that no single experiment, let alone a single interpretation of an experiment, could exhaust the relevant phenomena. Instead, the experimenter must resort to a multiplicity of experiments to bring the experimenter into close relation with Nature’s “inner mechanisms.” In their emphasis on situating the experiment, romantic experimenters adopted the standard of the object as a medium of reflexion in the manner of romantic art criticism. The essence of romantic art criticism was not the attempt for a definitive judgement of the work, but its consummation, completion, and systematization. The romantic criticism that begin with the Schlegel circle rejected the Enlightenment aestheticians’ and art critics’ method of weighing up a work of art against an inventory of aesthetic principles and casting a verdict upon it. Instead, they sought to assess the limits of the visible work and open it up to the realm of the invisible work, the idea of art. Confronted with the individual work, then, the challenge was to determine how the work individuated or incarnated the Thought of Art. Similarly, the aim of the romantic experiment was not to confer authority upon an experiment but to grasp its particular mode of individuation in its relation to the Totality or System. The problem of individuation, of the boundaries of experiments, demanded a reformulation of the relations between the piece and the totality, in short, the logic of the fragment.

4. EARLY ROMANTIC AESTHETICS As early as his prize-winning essay on literary aesthetics written at Copenhagen University in 1797 Ørsted championed ideas similar to those of Friedrich Schiller and other Germans who advocated aesthetics as the surest means to unite the sensible and spiritual powers of humans.27 While Kant and Fichte couched similar ideas in the more rigorous terms of the relations between Imagination and Reason, for Schiller the central value of the aesthetic was high pedagogy. “There is no other way of making sensuous man rational except by first making him aesthetic,” wrote Schiller in his Letters on the Aesthetic Education of Man (1795).28 The young intellectuals in Jena took this statement as a call to art. They also made much of the implication that there was no essential divide between literary and intellectual genres, especially between poetry and philosophy. As the 116th Athenäum fragment began, “Romantic poetry is a progressive universal poetry. Its goal is not merely to again unite all separated genres of poetry, and put poetry in contact with philosophy and rhetoric. It will and also should soon blend poetry and prose, brilliance and criticism, poetry of art and poetry of nature…”29 The task of romantic aesthetics as promulgated by the Schlegels et al. was to demonstrate the arbitrary and artificial nature of distinctions between genres, and to thereby recover the lost unity 27 28

29

Wilson, “The Unity of Physics and Poetry…” Friedrich Schiller, On the Aesthetic Education of Man (1795), translated by Elizabeth M. Wilkinson and L. A. Willoughby (Oxford: The Clarendon Press, 1982), p. 161. Friedrich Schlegel, “Athenäums-Fragmente,” in Kritische Schriften, ed. Wolfdietrich Rasch, second expanded edition (Munich: Carl Hanser, 1964), pp. 38–39.

224

R. M. BRAIN

of art and science, poetry and philosophy. As Friedrich Schlegel put it in his Kritische Fragmente (1797): “The entire history of modern poetry is a continuous commentary on the short text of philosophy: All art should become science, and all science should become art; poetry and philosophy should be united.”30 The possibility of overcoming the ordinary condition of separation was ensured by the extraordinary gift of philosophy, which after Kant enjoyed a renewed sense of its power to unite disparate categories, genres, and empirical sciences within a systematic unity. But for the romantics philosophy could not deliver on its promise by itself. Kant’s pursuit of reason’s systematic unity led inexorably to his antinomies of pure reason (e.g. the idea of the world as finite or infinite), the result of the infinitely progressive process which continually approximated but always failed to attain its goal. Schelling believed he had found a way out in the very “dialectical” character of this process, by postulating a development in the object itself as a counterpart to our rational understanding of it. But the romantics remained unconvinced that such a purely speculative answer escaped the inherent paradox. They argued that what was needed was a theory of the knowledge of objects—natural objects and works of art—which must be distinguished from knowledge of the System or the Absolute. Their view, in brief, was that the object, like everything real, falls within the medium of reflection, and therefore manifests the character of thinking. This is to say that the object itself possesses the capacity for self-knowing. “Everything that one can think, itself thinks—is a problem for thinking,” was the epigram that Friedrich Schlegel placed at the front of a collection of his late friend’s fragments that he edited. Novalis frequently emphasized that to observe a thing means only to arouse it to self-recognition; true experiment should be “the mere extension, division, diversification, augmentation of the object.”31 Everything that presents itself to human understanding as knowledge of an object is the reflex in the knower of the self-knowledge of that very being. All knowledge therefore forms an immanent connection in the absolute; it suspends the boundary that separates thing from knowing being, and makes the very term “object” both relative and provisional. Although this view may seem mystifyingly close to speculative philosophy, it had the effect of reorienting philosophy to a kind of empiricism, since thought could only realize itself through objects. The romantic enthusiasm for speculative philosophy thus retained an element of ambivalence or even resistance. Friedrich Schlegel expressed the problem in one of the Athenauem fragments, when he wrote, “It is just as deadly for the spirit to have a system or to not have one. Thus, he will have to [decide] how to join both.”32 Schlegel’s articulation of the problem of System versus nonsystem contained its answer within itself. Rather than seeking unity or the Absolute through general principles of reason, it would approach the System indirectly, cunningly, through its dialectical opposite, the fragment. This strategy depended on the very Schellingian assumption that the System was at

30 31 32

Friedrich Schlegel, “Kritische Fragmente,” in Kritische Schriften…, p. 22. Benjamin, Der Begriff der Kunstkritik…, p. 55. “Es ist gleich toedlich fur den Geist, ein System zu haben, und keins zu haben. Er wird sich also wohl entschleissen muessen, beides zu verbinden.” Friedrich Schlegel, “ ‘Athenäeums’ Fragmente,” in Kritische Schriften …, p. 31.

THE ROMANTIC EXPERIMENT AS FRAGMENT

225

work in the individual or particular, in miniature or microcosm, even if its precise modality was obscured by its finite character. In this way the romantics believed they had retained what Schlegel called “the spirit of system, which is something entirely different from a system.”33 Within this schema the object of cognition could be defined as a fragment, as a piece at once complete and torn away from a larger whole. From the start, then, the literary fragment served as the emblematic genre of the early romantic artwork.34 While the essential logic of the romantic fragment was inaugurated by the Jena romantics in the well-known fragments published in the Athenäum, Novalis’ Grains of Pollen, and Friedrich Schlegel’s Ideas, the idea of the fragment could be rendered in many ways other than simply the free-standing short written piece. Atheneum Fragment 77, for example, programmatically suggests that dialogue, letters, and certain forms of monument belong to the fragmentary. Even many of the romantics’ systematic expositions or “continuous” texts were composed along lines that expressed fragmentarity. There was a model for this, an Ur-text, of course. As Phillipe Lacoue-Labarthe and Jean-Luc Nancy observe, what were the Gospels but a garland of fragments which both individually and collectively glimpsed the figure of the Godhead?35 These considerations expressly left open the way for other artistic and scientific disciplines to develop their own properly fragmentary forms within their media of expression. Painters such as Phillip Otto Runge and Caspar David Friedrich famously used incompletion, motives involving ruins and monuments, and quotations from antique works of art to accomplish the aims of fragmentarity. Composers such as Robert Schumann, as Charles Rosen has splendidly shown, took the romantic fragment as the basis for a new emancipation of musical form. Schumann explored all of the possibilities of the Romantic fragment, composing works without a satisfactory beginning and end, playing with nonintegrated aspects of melody, building unstable, ambiguous, and even intelligible but inaudible elements into the music.36 Remarkably, Schumann’s technique derived some inspiration from a Ritter fragment where the physicist observed that “Tones are beings who understand each other, as we understand tone. Every chord may already be a mutual tone-understanding, and come to us as an already created unity.” Ritter’s fragment in turn drew upon Friedrich Schlegel’s fragmentary quip that “words often understand each other better than the people who use them.”37 The key to the romantic fragment therefore rested neither in the medium nor in the incompleteness of the work, but in its embodiment of the core logic of romanticism. After all, fragments of some kind had been around long before the 33 34

35 36

37

Friedrich Schlegel, Briefe an seinen Bruder, August Wilhelm (Berlin, 1890), p. 111. Phillip Lacoue-Labarthe and Jean-Luc Nancy, The Literary Absolute: The Theory of Literature in German Romanticism, trans. with an introduction and additional notes by Philip Barnard and Cheryl Lester (State University of New York Press,). My understanding of the romantic fragment has been substantially shaped by this text. Lacoue-Labarthe and Nancy, The Literary Absolute.…, p. 45. Charles Rosen, The Romantic Generation (Cambridge, MA: Harvard University Press, 1995), pp. 41–99. Rosen, The Romantic Generation. . , p. 70.

226

R. M. BRAIN

impudent youth of Jena. Succinct and incomplete literary and artistic forms— maxims, aphorisms, essays, pensées, remarks, opinions, anecdotes, etc.—enjoyed a long and flourishing history, of course; and were en vogue in the 18th century.38 The gates to the so-called Fragmentenflut in late 18th-century Germany were opened with Hamman’s Brokken or “crumbs”—an attack on Kantian architectonics through the assertion of the value of the unconnected and the unsystematic. Herder’s Ueber die neuere deutsche Literatur (1766), Klopstock’s Fragmente ueber Sprache und Dichtkunst (1779–1780), Lavater’s Physiognomische Fragmente (1775– 1778) instigated the so-called Fragmentenstreit of the 1780s. Some of these provided key resources for romantic uses of the fragment, notably Lavater’s work and Goethe’s Faust, first published as a fragment in 1790. Lavater slouched toward the romantics when he indicated that the very conception of his thought was fragmentary, taking the medium as the message. All of these authors vaunted the condensed or unfinished work, the detached piece, the residue of a broken ensemble, or the idea of fragmentary cogitation. But the romantics raised the ante on the genre, transfiguring the very idea of the fragment within the new horizon offered by philosophical idealism. Friedrich Schlegel, who gave first definition to the early romantic notion of the Fragment, insisted that the Fragment be a finished form. The well-known Athenaeum Fragment 206 offered this prickly definition: “A fragment should be like a little work or art, separated from the surrounding world and complete in itself like a hedgehog.”39 Ein Fragment muβ gleich einem kleinen Kunstwerke von der umgebenden Welt ganz abgesondert und in sich selbst vollendet sein wie ein Igel.

The hedgehog is a benign little creature, with a well-defined yet irregular form when it is frightened and rolled up into a little ball. Its quills invite us to ponder in what way it separates from the world—in romantic terms, “defines its finitude” (or in late romantic parlance, “individuates”). The hedgehog differs from the more aggressive porcupine in that it does not shoot its quills, yet it retains a capacity to wound those who attempt to breach its boundaries. Hence, as Charles Rosen observes, “the Romantic fragment draws blood only from those who handle it unthinkingly.”40 The romantic fragment demands that we attend to its specific character, especially the aggressive way it cuts itself off and assumes perspectives outside or distant from received understanding. This definition of the fragment alerts us not to mistake for true romantic fragments pieces that are merely incomplete or those which aim at fragmentation for their own sake. There are many “fragments”—some written by lesser writers, but some even by estimable romantic authors—that were hardly more than ordinary observations or run-of the-mill maxims, and thus failed to qualify as

38

39

40

Elizabeth Wanning Harries, The Unfinished Manner. Essays on the Fragment in the Later EighteenthCentury (Charlottesville and London: University Press of Virginia, 1994); J. A. Schmoll, Eisenwerth (ed.), Das Unvollendete als Kuenstlerische Form: Ein Symposium (Bern: Francke, 1959). Friedrich Schlegel, “Athenäums-Fragmente,” Friedrich Schlegel Kritische Schriften, ed. Wolfdietrich Rasch (Munich: Carl Hanser, 1970), p. 47. Charles Rosen, The Romantic Generation …, p. 48.

THE ROMANTIC EXPERIMENT AS FRAGMENT

227

true romantic fragments. Pretend or pseudofragments notwithstanding, the early romantic authors took particular interest in the way in which the world as one finds it comes broken in so many pieces, littered with shards, and gathered in ruins and heaps. If a merely broken piece did not qualify as a fragment, it nevertheless offered great potential if its accidental or involuntary character could be transfigured into a determinate and deliberate statement of fragmentation. Romantic authors (and later, composers) delighted in reinserting quotations from older sources in unlikely places where they seemed to generate naturally out of the new context. Robert Schumann, for example, would transform a musical joke into a tragic effect, or employ the prosaic or ungainly not simply for comic relief but seriously.41 Less dramatically, but no less effectively, romantic experimentalists advocated scavenging among the works of past authors for observations, techniques, and effects that might be reincorporated into the new science. Goethe’s historical treatise in the Farbenlehre pillaged 2,000 years of investigation and lore about color; for the rest he turned to painters, dyers, and other artisans with practical knowledge of the subject. Ørsted, in his 1805 lecture entitled “New Investigations into the Question: What is Chemistry?,” argued that a brief survey of the histories of metallurgy and alchemy suggested “how uncertain […] all external determinations of the boundaries of a science are…,” and that “so far we have only collected fragments, under various names, for this science, and that chemistry only constituted one of them.”42 Like the remains of fragmented Antiquity, the landscape of ruins, the archive of natural philosophy was littered with assorted facts gathered by all disciplines practicing chemistry, often in the name of other disciplines such as mining or Ørsted’s formative métier, pharmacy. Valuable fragments could be culled from adventitious observations arising from some of the most canonical experiments, such as those made by Ritter and Herschel regarding Newton’s prism experiments. What was required were new disciplinary histories to identify, and finally, recontextualize the stray, marginal, and alien facts and unacknowledged causes within a new series of experimentation.43 Disciplinary narratives had to be in constant state of revision, injected with the very same ceaselessness and flux manifested by Nature itself. The Romantic fragment referred to a process rather than a fixed state. In many instances of Novalis’ Pollen, or this one from Schlegel, for example, the Fragment carries the seeds of its own development. The Germans, they say, are the greatest people in the world in respect to the elevation of their sense of art and their scientific thought. Indeed: only there are very few Germans.44 Die Deutschen, sagt man, sind, was Höhe des Kunstsinns und des wissenschaftlichen Geistes betrifft, das erste Volk in der Welt. Gewiss: nur gibt es sehr wenige Deutschen.

Here the assertion of national pride barely sets in before it is abruptly (and wittily) undermined. The fragment thus sets in motion a process that remains incomplete.

41 42 43

44

Rosen, The Romantic Generation.…, p. 58. Ørsted, “What is Chemistry?…,” p. 198. On Ørsted’s practices of recontextualizing the work of others, see Kenneth L. Caneva, “Ørsted’s Presentation of Others’—and his Own—Work,” (in this volume). Friedrich Schlegel, “Kritische Fragmente,” in Kritische Schriften…, p. 22.

228

R. M. BRAIN

Yet, the fragment itself is formally complete, even if the content is incomplete, by implying a process still underway. It suggests the impossibility of capturing a dynamic, developmental process by reference: instability proves a more effective form of signification. Since classical antiquity maxims were presented as objects of contemplation, for ongoing interpretation. In this instance the romantic Fragment shows the act of interpretation already in motion, implying a past before the fragment begins and a future after it ends.45 Similarly, Schlegel writes, “One can become a philosopher, not be one. As soon as one believes he is one, he ceases becoming one.” [Man kann Philosoph werden, nicht es sein. Sobald man es zu sein glaubt, hört man auf es zu werden.] 46 Here the fragment adheres to the order of Becoming, of organic development, or more specifically, of autogeneration, which is the proper character of Selfhood.47 The instability of the Fragment, its delimited place in a larger state of Becoming, demands that it never appear alone but in the plural. The romantics never published a single fragment, but always collections of fragments. The individual fragment typically stood for itself and for the totality within which it participated; the totality was not the sum of the individual fragments but was glimpsed partially through the presence of each fragment. Ørsted, for example, frequently stipulated that his experiments represented infinite Nature in finite terms, such that “every well-executed investigation of a limited object reveals a portion of the eternal laws of the Whole.”48 But this does not mean that an experiment takes up a fraction of the whole (in a mathematical sense), and that the sum of the experiments would restore the whole. Rather, it replicates the Whole, or registers its copresence, in each fragmentary experiment. Each experiment, in other words, is a miniature—separate, individual, and complete in itself—that implies the world outside from which it is separated. As a methodological matter, Ørsted indicated that experiments be carried out in a series. If executed properly, the sweep of the series would allow reason to ascend to a comprehension of the Absolute. “Nature is infinite,” Ørsted wrote in “What is Chemistry?,” so we cannot construct it differently from the way we construct an infinite series, that is, by presenting a certain part of it and from this deducing the laws for the whole. In this way the construction of a finite part of nature can be extremely instructive. In experiments we force nature to make a construction or rather a reconstruction in front of our eyes, and what reconstruction could be more instructive for us than the one Nature itself shows us if only we know how to see it.49

For romantics, one might say, experiments never end, or rather finite experiments always end, but they entail a never-ending unfolding of further experimentation. This view was not akin to the reproach of experimenter’s regress (the charge that the justification of the experiment lies in arguments and judgements made outside of the performance of the experiment itself) since there was no claim that 45 46 47

48 49

Rosen, The Romantic Generation.…, p. 51. Friedrich Schlegel, “Athenäums-Fragmente,” in Kritische Schriften…, p. 31. Helmut Mueller-Sievers, Self-Generation: Biology, Philosophy, and Literature around 1800 (Stanford, CA: Stanford University Press, 1997). Ørsted, “Introduction to General Physics …,” p. 285. Ørsted, “What is Chemistry?…,” p. 199.

THE ROMANTIC EXPERIMENT AS FRAGMENT

229

experiments should bring about closure to an argument.50 On the contrary, the dialectical linkage of nature and reason ensures that the evolving character of both requires an “active empiricism,” to use Novalis’s term.51 Both Ørsted and Novalis agreed that the answer to what Schelling called nature’s ins Unendliche fliessende Funktion lay in experimental Serie or Reihen. Novalis called this the “view of an experiment as an intensifying series of points of view.” The serial unfolding of fragments enabled a shifting of scales from the miniature to the panoramic. While the individual fragment must be “characterized by a few strokes of the pen” (Atheneum fragment 302), the series of fragments amplified those minimalist strokes to the Infinite. The notion of intuition at work here derived from physiognomy—it was no coincidence that Lavater adopted the form of the fragment. But again, the romantics took the concept a step further. From those few strokes reason could be set in motion, and from the experimental object forces, laws, and the Godhead could be comprehended.

5. HOW TO MAKE AN EXPERIMENT LIKE A HEDGEHOG For the experimental natural philosophers who dipped in the same icy romantic waters, the task was to define their own version of the finite, fragmentary work. Novalis paved the way, asserting that what marked the genuine experiment was not a question put to nature but only observation fixed on the nascent self-knowledge of the object. Novalis defined the experimental adept as one for whom nature “reveals itself all the more completely through him, the more his constitution is in harmony with it.”52 In some experimenters this maxim led to a fascination for dowsing, clairvoyance, somnambulism, and other “sidereal” techniques which required an apparently harmonious relation to nature.53 It also led to a fascination with self-experimentation. Ritter famously took selfexperiment to an extreme, applying his galvanic prod to every sense organ and indeed every tissue of his body. He assembled his experimental experiments into his Fragmente aus dem Nachlass eines jungen Physikers, a collection of some 700 literary fragments concerning many branches of physics (dynamics, chemistry, acoustics, optics, galvanics, etc) as well as Heinrich’s philosophy, art, religion, and the polarity of the sexes. Ritter aimed the literary style of the fragments “somewhere between Novalis and Lichtenberg” in their embodiment of literary poles of naivité and sophistication, coyness and boldness, etc.54 The literary form of the fragment was best suited to express the principle, expressed by Ritter to Ørsted a few months

50

51

52 53

54

On the problem of experimenter’s regress see H.M. Collins, Changing Order: Replication and Induction in Scientific Practice (London and Beverly Hills: Sage, 1985). Fergus Henderson, “Novalis, Ritter, and Experiment. A Tradition of Active Empiricism,” in Elinor S. Shaffer (ed.), The Third Culture: Literature and Science (Berlin: 1998), pp. 153–170.. On Novalis’s notion of experiment see also Jürgen Daiber, Experimentalphysik des Geistes. Novalis und das romantische Experiment (Göttingen: Vandenhoeck & Ruprecht, 2001). Novalis [Friedrich von Hardenberg], cited in Walter Benjamin, Der Begriff der Kunstkritik.…, p. 54–55. Stuart Walker Strickland, “Circ*mscribing Science: Johann Wilhelm Ritter and the Physics of Sidereal Man,” (Ph.D. dissertation, Harvard University, 92). Ritter to Ørsted, Correspondance de H.C. Örsted…, vol. 2, pp. 228–229.

230

R. M. BRAIN

earlier that “as physicist one is a doubled nature,” inasmuch as the physicist both observes nature and is nature—a condition that was difficult to communicate to those lay people who had not developed this self-consciousness. When Ørsted confided to Ritter that he wanted to write a textbook of physics, Ritter remarked on how he wanted to do the same, but was unable to produce a contrived systematicity of the field or a correspondingly invented unity of self. “I would—almost—write it in aphorisms,” Ritter wrote, adding: “I fear the construed like death.”55 Given these remarks it is not surprising that the textbook was never written. Instead, Ritter imagined his Fragments as a means of inculcating this doubled self-consciousness in a young person “who wants or must become a physicist” by recapitulating, or at least emulating, this “literary autobiography.”56 Ørsted rejected self-experiment and did not publish collections of fragments, either. But he offered his own solution to the problem, conceiving and crafting his experimental art to suit the demands of the logic of fragmentarity. With one exception discussed below, Ørsted’s experiments were not composed as literary fragments, nor did he offer any theoretical discussion of the experimental fragment as such in any writings I am aware of. This was consistent with the romantic authors of fragments, of course, who strategically preferred to let fragments serve as definitions of fragments, and refrained from giving a pure theory of the work as fragment. Similarly, I shall argue, Ørsted’s form of experimentation illustrates the very idea of fragmentation in the very logic of its practice. To illustrate this logic, however, I will examine three specific examples from his writings. But first it is important to reiterate a few words about the experimental philosophy from which it departed. Let us restate Ørsted’s philosophy of experiment in a nutshell.57 Ørsted repeatedly emphasized that the limits of human knowledge derived from human finitude. Perception is the most obvious limitation: attempting to perceive the solar system with the telescope or the core of bodies by the cutting iron, or trying to fathom the indiscernible changes of nature’s incessant activity, we lose ourselves in the infinitely large, the infinitely small, the too-fleeting or the too-slow. “We also understand,” he wrote, “why our knowledge is only a weak reflection

55

56

57

“Ich würde– fast– in Aphorismen schreiben. Vor dem Construiren fürchte ich mich, wie vor dem Tod. Wenigstens stetze ich mich nie zum Construiren besonders hin. Ich muss die ächte Construction erst entdecken, dann reconstruire ich, u. stelle mich als Physiker freylich, als construirte sogleich.” Ritter to Ørsted, Correspondance de H.C. Örsted…, vol. 2, p. 109. Ritter to Ørsted, Correspondance de H.C. Örsted…, vol. 2, p. 228. For a similar view of the communication crisis brought on by romantic experiment see Simon Schaffer, “Self-Evidence,” Critical Inquiry 18 (Winter 1992), 327–362; and Stuart Walker Strickland, “The Ideology of Self-Knowledge and the Practice of Self-Experimentation,” Eighteenth-Century Studies 31:4 (1998), pp. 453–471. Ørsted made these sorts of remarks in many different writings, but I have drawn this account exclusively from his 1811 “First Introduction to General Physics,” which Ørsted described as an extended version of the introduction to his textbook published two years earlier. It was also published in Schweigger’s Journal für Chemie und Physik vol. 36 (Nuremberg, 1822), pp. 458–488, with an editorial comment that states “From this fragment, here offered for public evaluation, the readers will realize what and how much they can expect from this new work by one of the most brilliant … contemporary physicists…” See Ørsted, “First Introduction to General Physics. A Prospectus of Lectures on this Science,” in Jelved, Jackson, and Knudsen, Selected Scientific Works of Hans Christian Ørsted…, pp. 282–309.

THE ROMANTIC EXPERIMENT AS FRAGMENT

231

of the great Totality, for our reason, though related to the infinite in origin, is enveloped finitudes and can only break away from them conditionally. It is thus not given to any man to embrace the Whole.” But all is not lost, Ørsted reassured, the first step of holy experimentation is to reach the “happy realization” that the Whole can be attained through the shard, that “the few gleams that he is allowed to see are enough to raise him above the dust.” Ørsted emphasized that our capacity to know nature through reason stemmed from the truth that “our reason recognizes itself in objects.” The imprint of reason is found exhibited in natural phenomena as a cynosure to reach the higher laws of reason. It is a matter of moving from intuition to proper method: “every well-executed investigation of a limited object reveals a portion of the eternal laws the Whole.” It is a matter of delineating the precise configuration of laws at work in a given natural object, then following the trail of natural laws to their point of unity. At this point one arrives at the Idea of Nature (or thought of nature), which is none other than the revelation of the combined creative power and reason of the Godhead.58 Consider Ørsted’s well-known “Experiments on Acoustic Figures” (1810).59 Ørsted tells us that he has conducted a “series of many hundred experiments” to “get somewhat closer to the inner mechanism of these remarkable phenomena.” These were Chladni’s figures, of course, which produced innumerable phenomena not considered by Chladni that demanded amplification and reflection. Ørsted took great pains to describe his own method of producing the figures and how it differed from Chladni’s so that his trials could be reproduced: metal plates instead glass, supported at the edges instead of at the points of intersection, using iron filings instead of sand, and so on. He then described experiments in which the figures are generated in different ways. It appears initially that the figures are generated much in the way that Chladni would describe them—mechanically. Ørsted’s procedure up to this point is exacting and highly skilled, without any trace of philosophical thought, much like any paper that might be published in the Comptes Rendues of the Paris Academy of Sciences. But Ørsted then points out anomalous effects, especially the fact that the dust adheres more firmly to the plate in the acoustic figures than elsewhere. This phenomenon is further amplified until it becomes clear that a mechanical explanation is thoroughly destabilized. Ørsted then shows how other forces are at work, specifically weak electricities, aroused by the production of the tones, distributes a pattern of negative and positive charges on the surface of the plate, and is shared by the dust at each place covered on the plate. The apparent mechanical causes are shown to be illusory: it is the correlation of acoustic and electrical forces that generates the material form of the figures. At this point Ørsted has faced the finite object and identified both the special, perceptible (mechanical) forces at work within it, but also shown the activity of the imperceptible forces of electricity and chemistry involved in the phenomena. Ørsted seems to suggest that the interaction of forces indicates a unity of 58 59

H. C. Ørsted, “First Introduction to General Physics…,” p. 286. H. C. Ørsted, “Experiments on Acoustic Figures (1810),” in Jelved, Jackson, and Knudsen (eds.), Selected Scientific Works of Hans Christian Ørsted…, pp. 264–281.

232

R. M. BRAIN

natural laws. But this is hardly all—just at the moment where the experiment would appear to end, Ørsted begins the ascent to his real conclusions. Ørsted joins his results to those of Ritter on tones, who claimed to have shown how tones of high frequencies cease to become perceptible to the ear as sound, but become perceptible to the eye as light broken progressively into the colors of the spectrum. A “normal” synaesthesia, in other words, in which sense modalities displayed a similar correlation to that of physical forces. This point is crucial in Ørsted’s account because it shows the way we come to understand nature and our conception of it through a dialectical interplay. Both are explained through lawful reason, and the harmony of the explanations shows how Self and Nature are really one, the appearance of separation is now revealed to be as spurious as the distinction between natural forces. This implied the imperative—call it a kind of moral demand—that the experiment on the finite object was a means of bringing the Self into attunement with Nature. Through direct sensual contact between consciousness of nature and consciousness of self, the vital forces of nature and the human bodily self were brought into harmony. “All sensations spring from the same original force,” Ørsted concludes, adding that each sense registers an octave on a grand scale of sensation. The effects observed in the acoustic figures, in other words, revealed to the adept the lineaments of a comprehensive vision of natural laws. At this point Ørsted can no longer constrain his ecstasy, and he lets the Naturphilosophical chorus resound: What a great and profound and in itself necessary harmony, what mark of an allpervasive Reason. Here we see that it is not the mechanical sensory stimulation which pleases us in the tone, but the mark of an invisible Reason which lies in it. And now a flow of notes which pervades our whole being with joy. What profundity unknown to the listener is not hidden in a single chord, what infinite arithmetic in a whole symphony! And now, joined with this, the invisible forms which appear before our soul in obscure intimations while the note flow into the ear. In truth, we can repeat with joy and triumph at the nobility of our spiritual being that what fascinates and enraptures us in the art of music and makes us forget everything while our soul soars on the flow of notes is not the mechanical stimulation of tense nerves. It is the deep, infinite, incomprehensible Reason of Nature which speaks to us through the flow of notes.60

Through its internal workings, Ørsted’s experiment enables nature to manifest its laws and, through which we glimpse its lawfulness in the “mark of an invisible Reason.” The experiment thus serves as a miniature or a microcosm of the Absolute. It does so by figuring “in a few strokes” the absent Totality that is, nevertheless, essential to its working. However, Ørsted implies that an acquaintance with systematic Naturphilosophie might be as indispensable for following this movement of thought and grasping the sublime Reason of Nature as the technical skills necessary to produce the acoustic figures in the first place. Both experimental finesse and philosophical training were equally necessary for, as Ørsted frequently stressed, this kind of encounter with finite nature enabled the experimenter or artist to recognize himself in the outer world in a way that Naturphilosophie, perpetually confined to the circle of aprioristic systematics, could not. In this way the experiment, like the fragmentary artwork, completes the project of systematic philosophy. 60

Ørsted, “Experiments on Acoustic Figures,” …, pp. 280–281.

THE ROMANTIC EXPERIMENT AS FRAGMENT

233

It is not surprising that the experiments of Ritter (and perhaps Ørsted) inspired Romantic composers like Schumann to build into their works different kinds of absent or inaudible melodies: music never actually realized in sound but available to imagination of the listener (or more often to the musician reading the score). In the case of music, Charles Rosen has called attention to the way in which this aspect of the romantic fragment furthered a kind of esotericism available only to those initiated into the technical features of music, and perhaps also au courant with the theoretical games they might expect to find in a score.61 A similar problem arises with regard to the availability of this kind of experimental fragment to the public outside of the narrow circles of initiated romantic experimentalists. As we have seen above, Ritter regarded this as a problem, and in the end despaired that a method such as a public lecture could communicate the experience in form that the experimentalist grasped it. As we have seen, the plural character of romantic fragments always suggested a loose analogy between the relation of fragment to totality and a metaphysical or organic politics. Like Goethe (or Lavater) before them, the romantics advertised themselves as open and menschenfreundlich, terms opposed to dogmatic and authoritarian models. “To philosophize means to seek universal omniscience collectively,” [Philosophieren heisst die Allwissenheit gemeinschaftlich suchen] in Schlegel’s words.62 In proud contrast to the Newtonian networks or the French Academy, romantic experimentalists welcomed contributions from all sensitive souls, even if these, like Germans, turned out to be rather few in number. The romantic experimentalist remained troubled, as Strickland showed, because the experimental practice (among other romantic practices) tended toward an elitism at odds with its professed democratic aims. Ritter worried about the gap between performance and reception in the new mode of experimentation. He complained to Ørsted that he believed that even at its most persuasive and revealing the public lecture was too “self-contained” and prone to create an illusion of completeness for auditors, who simply remained ignorant of the historical, dialectical, and systematic elements that informed the experiment and which were held back as “the secret business of the teacher.”63 But Ørsted, ever the heir of Schiller’s Letters on the Aesthetic Education of Humanity, remained optimistic, believing that it was precisely this kind of sensible experiment, supplemented with a few well-placed strokes of philosophical reason, that could excite edifying effects in the layperson. Sensibility was here another way of indicating human finitude, to which even the most recondite philosopher had to own up. For Ørsted the self-containment of the public experiment formed its distinct advantage, for it forbade philosophical excess or shortcuts and required that experimenter and auditor alike attend to the precise manner in which the object separates from the world—its hedgehog-like character—and how its true place in the unity of Reason could be fathomed. Haryard University 61 62 63

Rosen, The Romantic Generation…, pp. 112–115. Schlegel, “Athenäums-Fragmente,” Kritische Schriften…, p. 67. Ritter, quoted in Strickland, “The Ideology of Self-Knowledge…,” p. 461.

ØRSTED AND THE RATIONAL UNCONSCIOUS LORRAINE DASTON

1. INTRODUCTION: THE UNCONSCIOUS RID OF THE ID From his earliest lectures on psychoanalysis in 1915–1917 to his very last works on the subject in 1938, Sigmund Freud remarked upon the vehement opposition provoked by the very idea of the unconscious, especially in scientific circles: “The concept of the unconscious has long been knocking at the gates of psychology and asking to be let in. Philosophy and literature have often toyed with it, but science could find no use for it.”1 From the outset, he insisted that the existence of the unconscious had been empirically proven, challenging “anyone in the world to give a more correct scientific account of this state of affairs, and if he does we will gladly renounce our hypothesis of unconscious mental processes. Till that happens, however, we will hold fast to the hypothesis; and if someone objects that here the unconscious is nothing real in a scientific sense, is a makeshift, une façon de parler, we can only shrug our shoulders resignedly and dismiss what he says as unintelligible.”2 This defiance in the face of scientific scepticism bordered on the defensive, and Freud felt compelled to explain collective resistance to the hypothesis of the unconscious, most famously in his account of the three blows science had dealt to human amour-propre: first Copernican astronomy displaced mankind from the center of the universe, then Darwinian evolutionary theory revealed the continuity of human and animal nature, and finally psychoanalysis dethroned reason and consciousness, showing that the ego “is not even master in its own house, but must content itself with scanty information of what is going on unconsciously in its mind.”3 Although in later theoretical works Freud emphasized that “even subtle and difficult intellectual operations” could be carried out in the unconscious, his more general identification of consciousness with the ego

1

2

3

Sigmund Freud, “Some Elementary Lessons in Psycho-analysis [1940],” in The Standard Edition of the Complete Psychological Works of Sigmund Freud, edited and translated by James Strachey, in collaboration with Anna Freud and assisted by Alix Strachey and Alan Tyson (London: The Hogarth Press and The Institute of Psycho-analysis, 1966–1974), vol. XXIII, p. 286. Sigmund Freud, Introductory Lectures on Psycho-Analysis [1916–1917] , in Standard Edition, vol. XVI, pp. 277–278. Ibid. vol. XVI, p. 285.

235 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 235–246. © 2007 Springer.

236

L. DASTON

and with “what may be called reason and common sense” and the unconscious with the id, instinct, and “the passions”4 has served to brand the unconscious as the repository of the irrational. Hence insofar as science lays claim to rationality, its alleged distrust of the unconscious is unsurprising. Yet in the century preceding Freud, other notions of the unconscious circulated, including some that located the highest faculties, rational and aesthetic, in that dark quadrant of the psyche. Hans Christian Ørsted’s “physics of the beautiful,” expounded in the form of a dialogue that featured musical and acoustical examples, developed a most un-Freudian conception of the rational unconscious. Although there is no evidence that Ørsted’s conception was more profound or influential than any of the other theories of the unconscious ventilated in 19th-century medical, philosophical, and literary sources,5 it combined elements of Naturphilosophie, Pythagorean music theory, experimental acoustics in the style of Ernst Florens Friedrich Chladni, and the post-Kantian aesthetic cult of genius into a pattern that endured long after each of the elements taken singly had been rejected. It is particularly striking to find this pattern of the rational unconscious in the acoustical writings of Hermann von Helmholtz, who notoriously rejected most of its components, especially Naturphilosophie. More generally, Ørsted’s rational unconscious strongly resembles late 19th- and early 20thcentury theories of creativity first developed in the context of art, especially musical composition, but then extended to science, especially mathematics. Before Hans Reichenbach, Karl Popper, Imré Lakatos, and other mid-20th-century philosophers of science pronounced scientific creativity a black box, defying all rational reconstruction, scientists themselves had embraced a view of the unconscious that was, pace Freud, at once creative and rational. Ørsted’s was among the earliest and oddest versions of this rational unconscious.

2. THE PHYSICS OF THE BEAUTIFUL Ørsted wrote a dialogue and several essays on the “physics of the beautiful” (and also on the physics of the “unbeautiful”); further remarks about the beauty of certain natural phenomena and experiments may be found scattered throughout his other scientific and philosophical writings, often recurring to the same examples, such as Chladni and Lichtenberg figures. His dialogues “On the Reasons for the Pleasure Evoked by Tones” and “The Natural Effects of Ordered Sounds” were apparently addressed to a popular audience; the essays on “The Natural Theory of the Beautiful” and “On the ‘Unbeautiful’ in Nature in its Relation to the Harmonic Beauty of the Whole” were originally written for the Royal Danish Society of Sciences in Copenhagen.6 However, there is 4 5

6

Sigmund Freud, The Ego and the Id [1923], in Standard Edition, vol. XIX, pp. 25–26. For general surveys of these theories, see Henri F. Ellenberger, The Discovery of the Unconscious: The History and Evolution of Dynamic Psychiatry (New York: Basic Books, 1970), pp. 182–330; Alexander Jacob, De Natura Naturae: A Study of Idealistic Conceptions of Nature and the Unconscious (Stuttgart: Franz Steiner Verlag, 1992), pp. 56–158. These papers were probably presented to the Royal Danish Society circa 1807: Ole Immanuel Franksen, H. C. Ørsted: A Man of the Two Cultures (Birkerød: Strandbergs Forlag, 1981), p. 14.

ØRSTED AND THE RATIONAL UNCONSCIOUS

237

considerable overlap in content and approach among these essays, and I shall treat them as a whole.7 In Ørsted’s dialogues on the natural foundations of the aesthetic response to music, five friends debate the reasons for music’s effects on the human senses and psyche, each character representing a theory: Felix defends the Romantic position of ineffable sublimity, Waldemar is the sensualist who ascribes musical pleasure to the corporeal organization of the nerves, Hermann maintains that it stems from the exercise of reason, Julius insists on the role of the imagination, and finally Alfred argues for an unconscious but rational enjoyment based upon the detection of regularities and symmetries. In the second dialogue, Felix, Waldemar, Alfred, Hermann, and Hermann’s wife Sophie (Julius having disqualified himself from further participation by not attending closely enough to Alfred’s arguments in the first dialogue) meet again 25 years later to discuss the effects of music and rhythm on the body as well as the soul. It might be noted in passing that Alfred, qua natural philosopher and mathematician, demands and for the most part commands the deferential respect of the other interlocutors: when Julius is not prepared to devote the necessary time and attention to the theoretical, experimental, and mathematical study of the “physics of tones,” he is summarily dismissed8; in the second dialogue, the self-deprecating and earnest Sophie becomes Alfred’s principal auditor, after assuring him that she would be unworthy of his instruction should she find his explanations too dry and detailed—thus taking her place in the long line of intelligent but professedly ignorant women in scientific dialogues stretching from the Marquise in Bernard de Fontenelle’s Entretiens sur la pluralité des mondes (1686) to Margharete Bohr in Michael Frayn’s recent play Copenhagen (1998). Since Alfred dominates both dialogues and advances positions that echo those set forth in Ørsted’s other essays on the subject, it may be fairly assumed that he speaks for Ørsted. Alfred makes his first appearance dreamily drawing figures in the sand, an activity reminiscent of the geometric parts of Plato’s dialogue Meno, in which a slave boy is made to prove the Pythagorean theorem under Socrates’ directed questioning. Just as the slave boy was unconscious of his knowledge of mathematics before his conversation with Socrates, so appreciative listeners are unconscious of the hidden symmetries and natural laws they in fact discern in music, or so Alfred will argue. But the figure Alfred is tracing in the sand is not a right triangle; it is instead a figure from experimental acoustics: “I was trying to draw a few tones.”9 What Alfred has in mind were the recent and much-publicized discoveries by the German physicist Chladni that when a glass or metal plate strewn with sand is stroked with a violin bow, regular figures suddenly are formed in the sand, 7

8

9

I shall rely on the German translations that first appeared in 1851 in a collection entitled Neue Beiträge zu dem Geist in der Natur; I very much regret that my lack of Danish prevents me from working from the original versions. Hans Christian Ørsted, Ueber die Gründe des Vergnügens, welches die Töne hervorbringen, in Neue Beiträge zu dem Geist der Natur, trans. K. L. Kannegießer , 2 vols., 2nd ed. (Leipzig: Carl B. Lorch, 1855), vol. 1, pp. 1–38, on p. 34; Ørsted, Die Naturwirkung des geordneten Lautausdrucks, in ibid. vol. 1, pp. 39–67, on p. 45. Ørsted, Ueber die Gründe, p. 15.

238

L. DASTON

making patterns of vibration in the plate visible.10 Chladni had been inspired by Georg Christoph Lichtenberg’s experiments on electrical figures, in which an electrical spark formed characteristic figures in powder strewn over a non-conducting plate according to whether subjected to positive or negative electrical charge.11 Ørsted performed experiments on both Chladni and Lichtenberg figures using a finer powder than sand, the seeds of club moss, which allowed him to observe the processes by which the figures coalesced more closely: “Mr. Chladni’s experiments are astonishing to anyone who sees them for the first time on account of the regularity of the figures which are produced by a single stroke of the bow, as if by magic. My experiments do not have the same charm, but perhaps they are more instructive because of the comparative slowness of my process, which allows for their effects to be studied more advantageously.”12 Ørsted had Alfred echo this theme of using the Chladni figures to make “the whole inner mechanism visible” by which perception registers and judgment appreciates musical tones.13 Throughout Ørsted’s works, analogies between visible forms, for example between Lichtenberg figures and organic forms, were repeatedly remarked upon as clues to deeper truths: “These symbols undoubtedly deserve our fullest attention, for they reappear everywhere, and who knows whether all of Nature’s mathematics does not lie hidden in them!”14 It was in these forms—the circles cast by a stone in water, the parabola traced by a fountain, the hyperbolas in sand made by vibrations of a metal surface—that we find “what our thought created once again in nature; what were thoughts in us, stand outside us as natural laws.”15 In the context of the aesthetic enjoyment of music, Chladni figures and other regular forms seemed to Ørsted to provide the basis for reconciling the sensual and rational aspects of acoustical pleasure. Using the example of the circle, Alfred attempts to persuade his friends that the innumerable symmetries of this figure, some immediately obvious to the eye, others discernable only through mathematical inquiry, fuse together into a unity that strikes the onlooker as beautiful: “You regard this infinite unity in infinite mutation with admiration, without becoming conscious of the whole magnitude of the idea.”16 This “idea” is a “unification of

10

11

12

13

14

15

16

Ernst Florens Friedrich Chladni, Entdeckungen über die Theorie des Klanges (Leipzig: Weidmanns Erben und Reich, 1787). Franz Melde, Chladni’s Leben und Wirken, 2nd ed. (Marburg: N.G. Elwert’sche Verlagsbuchhandlung, 1888), pp. 11–12. Chladni also described Lichtenberg as the Geburtshilfer of his later hypothesis that meteorites were of extraterrestrial origin: ibid., pp. 48–50. Hans Christian Ørsted, “A Letter from Mr.Ørsted, Professor of Philosophy in Copenhagen, to Professor Pictet on Acoustic Vibrations [1805],” in Karen Jelved, Andrew D. Jackson, and Ole Knudsen, trans. and eds., Selected Scientific Works of Hans Christian Ørsted, with an Introduction by Andrew D. Wilson (Princeton, NJ: Princeton University Press, 1998), pp. 181–184, on p. 182. Ørsted, Ueber die Gründe, p. 15. Lichtenberg also seized upon this aspect of Chladni’s discovery: in a 1792 letter introducing Chladni to the astronomer Heinrich Olbers, he praised what “this excellent man [Chladni] has done for the theory of the vibrations of sounding bodies by making these visible [durch Sichtbarmachung derselben].” Quoted in Dieter Ullmann, Chladni und die Entwicklung der Akustik von 1750–1860 (Basel/Boston/Berlin: Birkhäuser, 1996), p. 49. Hans Christian Ørsted, “On the Harmony between Electrical Figures and Organic Forms [1805],” in Jelved et al. Selected Scientific Works, pp. 185–191, on p. 185. Hans Christian Ørsted, Zwei Kapitel der Naturlehre des Schönen, in Neue Beiträge, vol. 1, pp. 69–125, on p. 78. Ørsted, Ueber die Gründe, p. 19.

ØRSTED AND THE RATIONAL UNCONSCIOUS

239

reason and sense,” which is accessible to the inner faculty of intuition (Anschauung) without the intermediary of conscious thought.17 When Alfred’s friends protest that this unconscious quality of the aesthetic idea robs it of its rational character, he replies that the thoughts and the detection of relationships among them that have been won by conscious effort gradually sink into unconsciousness, where they blend into a genuine whole, the “idea.” The recognition of the beautiful is the impression of such ideas, of which we have ceased to be conscious. This process of the unconscious forging of ideas “we call a mental intuition [geistige Anschauung]. The view of the beautiful, insofar as it is unmixed with anything else, proceeds from the idea without [conscious] knowledge, although the pleasure experienced stems from the secret accord between sensuous nature and reason.”18 Wherever regularity exists, this “secret accord” can be found; even crystals formed by chance bear “the stamp of reason.” Wherever natural laws are at work, exalts Alfred, “I deeply feel in them an infinite, unfathomable Reason, of which I can comprehend piecemeal only an incalculably small part. In short, nature is for me the revelation of an infinitely alive and active Reason.”19 This equation of natural laws with the laws of reason repeats like a leitmotif in Ørsted’s writings,20 and has been persuasively traced to his admiring but not uncritical reading of Schelling and other Naturphilosophen.21 However, there were also significant divergences between Ørsted’s and Schelling’s versions of Naturphilosophie, particularly on the status and role of the unconscious. For Schelling, the task of a transcendental philosophy of nature is to equate the “unconscious [bewußtlose], or as it is also called, the real activity” of nature with the ideal, i.e. those forms of intelligence which are “consciously productive [mit Bewußtsein produktiv].” Hence the regularities of nature, including “the sublime geometry” of celestial mechanics, are to be explained not by claiming that nature is the most perfect realization of geometry, but rather by comprehending that “the most perfect geometry is that which produces nature, by means of which explanatory form the real is transposed into the ideal.” Since nature is nothing other than the visible incarnation of human understanding, it must necessarily produce regularities. In this scheme, nature represents the unconscious and

17

18 19 20

21

Ibid. p. 21. Cp. Kant’s remarks on the “infinitely many splendid properties” of the circle, and the delight it and other mathematical figures inspire by revealing “intellectual purposiveness” in nature: Immanuel Kant, Critique of Judgment [1790], translated by Werner S. Pluhar (Indianapolis and Cambridge: Hackett, 1987), II. 62, pp. 239–240 (Akademie ed. pp. 362–364). Ibid. p. 23. Ibid. p. 27. Cp. Hans Christian Ørsted, “View of the Chemical Laws of Nature Obtained through Recent Discoveries [1812],” in Jelved et al., Selected Scientific Works, pp. 310–392, on p. 384: “However, what finally gives the study of nature its ultimate meaning is the clear understanding that natural laws are identical with the laws of reason, so they are in their application like thoughts; the totality of the laws of an objcet, regarded as its essence, is therefore an idea of Nature, and the law or the essence of the universe is the quintessence of all ideas, identical with absolute reason. And so we see all of nature as the manifestation of one infinite force and one infinite reason united, as the revelation of God.” Andrew D. Wilson, “Introduction,” in Jelved et al., Selected Scientific Works, p. xxxvi; Timothy Shanahan, “Kant, Naturphilosophie, and Ørsted’s Discovery of Electromagnetism: A Reassessment,” Studies in History and Philosophy of Science 20(1989): 287–305.

240

L. DASTON

what Schelling called “the real,” mind the conscious and “the ideal.” Nature is redeemed by Naturphilosophie by becoming “the organ of self-consciousness”; all necessity in nature points in the direction of this self-consciousness.22 Ørsted tried his hand at this kind of transcendental Naturphilosophie in an early work obviously indebted to Kant and Schelling, but significantly shifted the emphasis from self-consciousness to perception as the departure point for a deduction of natural laws: “…we shall try to deduce general and strictly necessary laws of Nature from the nature of perception itself.” In the same work, he praised Schelling’s Idee zu einer Philosophie der Natur (1797) and Von der Weltseele (1798) for the “beautiful and grand ideas” expressed therein, but complained that Schelling seemed incapable of distinguishing the empirical from the a priori, or, within the empirical, even true from false.23 Although Ørsted remained faithful to Schelling’s identification of the laws of nature with the laws of reason, he displaced the unconscious from nature to mind and elevated its status from primordial condition to telos. The very highest exercises of reason were performed unconsciously; moreover, they became unconscious through the training of the body and the senses. The unconscious “inner sense” responsible for the feeling for the beautiful could, Ørsted thought, be developed and refined through practice. After all, he pointed out, people whose work requires that they constantly make measurements or calculations eventually develop the knack of judging proportions or reckoning sums without laborious, step-by-step thought: “What was earlier thought out has become something immediate for them.” Knowledge is originally won through the deliberate activities of thought and sense, but becomes literally incorporated [einverleibt] and thereby participates in the “immediacy of intuition [Anschauung].” Because this “inner sense” was a genuine sense, like sight or hearing or touch, it necessarily contained a bodily as well as a mental component. The “thought-pictures” of intuition “belonged not merely to reason, but also to the senses.”24 The effects of music also worked upon the body, as well as upon reason. Just as one plucked string could set another in motion through sympathetic vibrations, so, explained Alfred to Sophie, could music work upon the nerves of the human body. Chorales were particularly well-suited to bring “the restless interior to peace and order”25; people walking in indeterminate tempo will also have their gait “ordered” by the sound of drumbeats.26 Less rhythmic, more tonal music worked more on mood than on movement. Mathematical music theory and physical acoustics might investigate and calculate these effects, but such conscious deliberations played no part in either the creativity of the composer or the pleasures of the listeners: “This is unconscious.”27 Knowledge becomes unconscious, and hence immediate, through the formation of habits, corporeal and mental. 22

23

24

25 26 27

F.W. J. Schelling, Einleitung zu seinen Entwurf eines Systems der Naturphilosophie [1799], edited by Wilhelm G. Jacobs (Stuttgart: Reclam, 1988), pp. 19–21. Hans Christian Ørsted, “Fundamentals of the Metaphysics of Nature Partly According to a New Plan [1799],” in Jelved et al. Selected Scientific Works, pp. 46–78, on pp. 47, 77. Hans Christian Ørsted, Ueber das “Unschöne” in der Natur, in Neue Beiträge, vol. 1, pp. 127–142, pp. 130–132. Ørsted, Naturwirkung, p. 57. Ørsted, Zwei Kapitel, p. 86. Ørsted, Ueber die Gründe, p. 32.

ØRSTED AND THE RATIONAL UNCONSCIOUS

241

Much of this doctrine of the aesthetic unconscious is reminiscent of Heinrich von Kleist’s famous essay Über das Marionettentheater (1810), in which the graceful but wholly mechanical dance of marionettes inspires reflections on the “disorder in the natural gracefulness of human beings wrought by consciousness.” A boy loses his effortless grace by conscious study of his own movements in the mirror; conversely, a brute bear outfences a master swordsman. In these examples, the unconscious governs bodily aptitude, not aesthetic judgment. Reason and knowledge appear only at the end of the essay, where Kleist speculates that only “an infinite consciousness” can regain the grace otherwise achieved in the organic world only when “reflection becomes darker and weaker”: “Must we once again eat from the Tree of Knowledge, in order to fall back into the state of innocence? Yes indeed, he replied; that is the last chapter of the history of the world.”28 I do not wish to draw a line of direct influence between Kleist and Ørsted in either direction, only to point out the similarity of the assumptions concerning the unconscious as the source of unforced beauty, and the uneasy proximity between this view and the reduction of art to the mechanical dance of the marionettes, graceful because soulless. Ørsted seems to have been aware of this dangerous proximity. After Alfred has triumphantly proclaimed the musical creativity of even a Mozart as the work of the unconscious, Felix exclaims in horror: “But then the artist would be a machine.” Alfred hastens to reassure him that since nature is reason revealed, a fortiori mind is as well, and the genial mind is nothing less than a “spark of divinity.” The artist “through a felicitous feeling discovers and makes what many people in many years could not plumb with their understanding.”29 Alfred even goes so far as to revive the much-maligned term Enthusiasmus in its original sense of divine afflatus to describe the artistic creativity, which is rooted in “a deep reason, that no finite understanding may comprehend… With reverence then shall each and everyone honor art, which itself knows how to honor nature and reason.”30 In short, Ørsted’s response to the uncomfortable association between the unconscious and the mechanical was similar to that of Kleist: the creative unconscious becomes the infinite divine consciousness; the genius is he through whom the godhead—and reason—speak. Ørsted has miniaturized and internalized Schelling’s nature, which was unconscious but nonetheless embodied reason: it is now the human unconscious, which recognizes and creates beauty in the image of reason, and, in its most genial forms, in the image of God.

3. THE PLEASURES OF THE UNCONSCIOUS Although Ørsted’s blend of Naturphilosophie and natural theology was resolutely purged, at least in Germany, by the generation of scientists that came to maturity around 1848, his vision of the rational unconscious, now shorn of transcendental 28

29 30

Heinrich von Kleist, Über das Marionettentheater [1810] in Der Zweikampf und andere Prosa, ed. Christine Ruhrberg (Stuttgart: Reclam, 1984), pp. 84–92, on pp. 89, 92. Ørsted, Ueber die Gründe, p. 32. Ibid. pp. 37–38.

242

L. DASTON

philosophy and the divine afflatus, lingered on in the most influential work on musical acoustics of the latter half of the 19th century: Helmholtz’s Die Lehre von den Tonempfindungen (1863). From the standpoint of physics, physiology, mathematics, and musicology, Helmholtz’s treatment of “the physiological foundation of music theory” was vastly more ambitious and accomplished than Ørsted’s tentative essays. Yet in the final chapters on “Relationships to Aesthetics,” Helmholtz also turned to the rational unconscious to explain the sense of musical beauty. In contrast to Ørsted, who believed that certain geometric figures, certain chords, and certain light effects were indubitably and universally experienced as more beautiful than others by all healthy human observers,31 Helmholtz acknowledged that the history of musical composition showed that scales and harmonic structures were the product of artistic invention, not the natural construction of the ear. However, just as many architectural styles could be constructed out of the same building blocks, so, asserted Helmholtz, all musical traditions necessarily had recourse to the same physiological elements. But these elements, though rational, were not present to the conscious understanding: That beauty is bound by laws and rules, which depend on the nature of human reason, is at present no longer doubted by anyone who has himself thought about aesthetic questions or studied recent aesthetic works. The difficulty is that these laws and rules, upon which beauty depends and according to which it must be judged, are not given by the conscious understanding, and neither the artist, while he creates the work, nor the spectator or auditor, while he enjoys the work, are conscious of them.…[Art] must create as imagination envisions, lawfully without a conscious law, purposefully without a conscious purpose.32

Not only were these rational laws and rules not accessible to consciousness; to detect conscious deliberation in a work of art was fatal to aesthetic pleasure, as fatal as Kleist had found conscious deliberation to be for bodily grace. Yet paradoxically, Helmholtz also believed that aesthetic enjoyment could be intensified through the discovery of “the purposefulness, the coherence, and the balance of all the separate parts” of a musical composition or painting, although such probing analysis was not essential to aesthetic pleasure. Helmholtz was convinced that the pleasure of the beautiful depended on the consonance between the work of art and the nature of the healthy human mind, but how could this “lawlikeness” be registered immediately by inuition [Anschauung], without entering consciousness? Nor was the unconscious nature of aesthetic judgment incidental to pleasure:“This unconsciousness of the lawlikeness [diese Bewusstlosigkeit des Gesetzmässigen] appears not as an incidental matter to the effects of the beautiful on our minds, that can or cannot be the case, but it is evidently rather the principal thing, the chief point.”33 It is exactly the experience of sensing the order and lawlikeness of a work of art without being able consciously to analyze it that is the essence of aesthetic

31

32

33

According to Ørsted, the circle was more beautiful than the triangle, the major chord composed of ground tone, third, and fifth than the minor chord, light than darkness: Ørsted, Ueber die Gründe, pp. 23, 31; Zwei Kapitel, p. 99. Hermann von Helmholtz, Die Lehre von den Tonempfindungen (Braunschweig: Friedrich Vieweg und Sohn, 1863), pp. 552–553. Ibid. p. 554.

ØRSTED AND THE RATIONAL UNCONSCIOUS

243

enjoyment. Just as the creative artist is unconscious of the processes that bring forth a work of genius, and which would demand indefinite “time, deliberation, and work” to reconstruct consciously, we the auditors or spectators are unconscious of the processes that allow us to appreciate its beauty, our “feeling of tact and taste” paralleling that of the artist in the act of creation. For Helmholtz it was essential that both processes, creation and appreciation, be rational, though unconscious. For only through this rational element does aesthetic experience gain a moral dimension, the “moral elevation and the feeling of beatific satisfaction, which is called forth by immersion in genuine and high works of art.” Even in this plunge into the darkest depths of the human mind, which are opaque to conscious inspection, the art lover discovers an “order of the world, which is governed by law and reason in all of its parts,” and which is the basis for “trust in the healthy primordial nature [gesunde Urnatur] of the human mind.”34 But here as well the sine qua non of moral edification as well as of aesthetic enjoyment was the unconsciousness of the processes involved. Only those parts of a work of art that resisted attempts at complete analysis could provide pleasure and inspiration. It was exactly the elusive quality of this hidden rationality that certified the greatness of a work of art. In an example that is as eerily prescient of the later philosophy of Ludwig Wittgenstein as Helmholtz was eerily reminiscent of Ørsted’s Naturphilosophie, Helmholtz discussed how obvious family resemblances (say the same nose in father and daughter) kindle little aesthetic interest, but the more ineffable the resemblance, metaphorical (e.g. between musical intervals) as well as literal (e.g. two faces in a painting), the greater our fascination and admiration. It is the impenetrable, unconscious quality of these affinities that cause pleasure; the very same relations spelled out consciously and conscientiously by waking reason would be dull and lackluster—one is tempted to add, wissenschaftlich. Of course the unconscious also played a central role in Helmholtz’s theory of perception: the unconscious inductive inferences that are built up through experience by association. These inferences have the same logical character as conscious inductive inferences of the sort dealt with by John Stuart Mill; however, if they are in error, as in the case of sensory illusions, they cannot be corrected by better conscious knowedge, because “the induction is formed by an unconscious and involuntary action of memory, which appears to our consciousness as an alien, irresistable force of nature.”35 Helmholtz ascribed all illusion, including the illusions of the stage, to such rash, unconscious inductions, which can be intellectually (but not perceptually) corrected by conscious deliberation. Instruction in Copernican astronomy may teach us that the earth revolves around the sun, but we still see the sun sink below the horizon every day. Here the unconscious does not appear in the same rosy light as in Helmholtz’s aesthetic speculations: although there is nothing irrational about inductive inferences per se, unconscious inductive inferences tend to be premature and superficial, and extremely difficult if not impossible to correct by conscious consideration, which knows better. The unconscious is once again 34 35

Ibid. p. 555. Hermann von Helmholtz, Handbuch der physiologischen Optik (Leipzig: Leopold Voss, 1867), p. 450.

244

L. DASTON

inaccessible to conscious scrutiny, but in the context of perception, this is a source of error and illusion, not aesthetic pleasure and moral edification. A more searching light is thrown upon Helmholtz’s notions of the rational unconscious and aesthetic experience by what may appear at first to be a more remote source: his writings on the differences between art and science, especially his discussions of Goethe as natural scientist. Goethe had, Helmholtz claimed in an 1853 lecture, understood nature as a work of art—a possibly fruitful view in the youthful science of natural history, but fatal in the mature science of optics. By insisting in the Farbenlehre that optics not abandon the “realm of sensory perception,” Goethe had fundamentally misunderstood the abstract, conceptual nature of physical science.36 In a considerably later lecture on Goethe’s science, Helmholtz conceded that art and science might share a thunderbolt flash of inspiration, “which must precede all casting into words,” but thereafter their paths diverged sharply, the one into conscious concepts expressed in words and the other into the “immediate mental intuition [geistige Anschauung], the aroused emotion, of which the poet himself is barely aware.”37 In yet another lecture on Goethe as scientist delivered in 1862, Helmholtz deepened the contrast between the unconscious “tact” that guided artistic inductions about human character, and the “sharply explicit, general rules and laws” characteristic of the sciences. No great poet, painter, or composer worked “consciously with general rules and abstractions”; yet this was the core of scientific thinking. Art welled up spontaneously from the depths of unconscious memory; science was the “conscious logical activity of our mind in its purest and most perfect form,” accompanied by the unspontaneous virtues of painstaking care, exactitude, and caution.38 Science was “the iron work of self-conscious inference [that] demanded great perseverance and caution [Hartnäckigkeit und Vorsicht]”39; art was intuition, “the opposite of thinking, that is of conscious comparisons…[and] without mental effort.”40 Science was conscious, even self-conscious, and arduous; art was unconscious and spontaneous, an effortless outpouring. Despite the striking similarities between Helmholtz’s and Ørsted’s acccounts of the role of the rational unconscious in aesthetic appreciation, especially in the realm of acoustics and music theory, this opposition between dour, hardworking science and exuberant art would probably have astonished Ørsted. He took it for granted that the beautiful forms of art and science, the geometry of the circle and the structure of the chorale, were governed by the same rationality and appreciated with the same spontaneous pleasure. He unhesitatingly described Chladni figures as “beautiful,” and thought that the “joy [Freude] with which naturalists 36

37

38

39 40

Hermann von Helmholtz, Über Goethe’s naturwissenschaftliche Arbeiten [1853], in Vorträge und Reden, 5th ed., 2 vols. (Braunschweig: Friedrich Vieweg und Sohn, 1903), vol. 1, pp. 23–45, on pp. 38–41. Hermann von Helmholtz, Goethes’s Vorahnungnen kommender naturwissenschaftlicher Ideen [1892], in Vorträge, vol. 2, pp. 335–361, on p. 344. Hermann von Helmholtz, Über das Verhältnis der Naturwissenschaften zur Gesammtheit der Wissenschaft [1862], in Vorträge, vol. 1, pp. 158–185, on pp. 172–176. Ibid. p. 178. Helmholtz, Goethe’s Vorahnungen, p. 341.

ØRSTED AND THE RATIONAL UNCONSCIOUS

245

consider these objects [i.e. regular forms like crystals] testify sufficiently to their beauty.”41 Helmholtz however sought the secret of aesthetic pleasure precisely in those unconscious aspects of creativity and appreciation which, though rational, came without the sweated labor of conscious deliberation. And conscious deliberation—careful, painstaking, exact—had become in his view the essence of science. This is why the conscious analysis of the underlying rationality of a work of art might be described as wissenschaftlich, worthy but joyless. Without yet indulging in the forbidden passions of the Freudian id, the rational unconscious still managed to have all the fun.

4. CONCLUSION: CREATIVITY AND ITS DISCONTENTS It is a piquant irony to find a transmogrified version of Schelling’s Naturphilosophie nestled within a weighty scientific treatise by Helmholtz. To be sure, there are substantial differences: Schelling’s unconscious nature has become the rational unconscious of the human mind; all talk of the infinite absolute and the necessary determination of nature by mind has vanished, as has the godhead. What remains is the conviction that beauty in art is created and appreciated by the unconscious apprehension of rational laws. Without the middle term of Ørsted’s “physics of the beautiful,” these resemblances might remain too faint to be discerned. Although Helmholtz shed Ørsted’s conviction that the laws of nature are identical to those of reason, and that nature was nothing less than a revelation of the divine, he clung to the idea of a rational unconscious that can plumb the hidden regularities embedded in a work of art, especially in music, by direct intuition, without the lumbering mediations of conscious thought. This conviction is all the more striking for the contrast it poses to Helmholtz’s other, more disparaging remarks on the role of the unconscious in producing and perpetuating perceptual illusions. There is a scarlet thread that runs through all these versions of the rational unconscious, from Schelling to Ørsted to Helmholtz and beyond, and that is the creativity of the genius. It was the genius who tapped the deepest roots of the rational unconscious, and whose own rational unconscious flowed indistinguishably into that of nature and even God. Kant had notoriously denied science any claim to genius, since genius could neither be taught nor learned: one might master the profound insights of Newton’s Principia by long study, but no amount of study would produce a Homer (or even a Wieland), according to Kant. Talent proceeded by the mastery of rules, but it is through genius that “nature gives the rule to art.”42 True to this maxim and Kant’s restriction of genius to art, the accounts of the rational unconscious by scientists like Ørsted and Helmholtz focus on artistic creativity, even though they identified that creativity with the eminently scientific undertaking of grasping the laws of nature and reason. In principle, a Laplacean demon might be able, with infinite time and calculational power at its

41 42

Ørsted, Ueber die Gründe, p. 26. Kant, Critique of Judgement , II.46–47, pp. 174–178 (Akademie ed. pp. 307–310).

246

L. DASTON

disposal, to dissect a Mozart symphony into its rational parts, but both Ørsted and Helmholtz were firm in their conviction that this monumental exercise of the conscious understanding would neither explain the darkling processes by which Mozart had composed the symphony, nor illuminate those by which listeners took pleasure in it. The demon would understand, but not appreciate the symphony. And since Helmholtz at least believed that science must be submitted to the clear light of consciousness, this seemed to preclude any application of the rational unconscious to scientific creativity, except at the very first instant of inspiration. However, later students of scientific creativity flirted with the rational unconscious. Jacques Hadamard’s study of mathematical creativity featured stories of problems solved unconsciously, after the mathematician had put aside conscious attempts in frustration and despair. At a moment of complete banality—Henri Poincaré stepping onto a bus—the long sought-after answer would burst into consciousness fullblown, like Athena emerging from the head of Zeus.43 There could be no question of any contribution made by conscious deliberation, except by laying the groundwork that was subsequently appropriated and built upon by the unconscious. Similarly, the novelist Arthur Koestler portrayed Johannes Kepler as “sleep-walking” toward the solution to the vexed problem of the Martian orbit, his unconscious drawing ellipses while his consciousness still hewed to circles.44 In these cases there could be no question that, although its workings were mysterious, the unconscious was rational, indeed hyperrational, capable of insights that wholly escaped the conscious mind, just as works of artistic genius could not spring from conscious effort. For the most part, however, scientific creativity has been a tabu topic in the history and philosophy of science. Even those who have rejected Hans Reichenbach’s distinction between the contexts of discovery and justification still regard anything associated with the murky unconscious as ipso facto irrational, and therefore unlikely to have much to do with the finished products of science. Indeed, the Manichean logic of the long-lived distinction between discovery and justification has probably reinforced the perceived irrationality of the unconscious, by opposing it diametrically to the alleged rationality of public scientific discourse. The debates that raged over the rationality or irrationality of scientific revolutions in the wake of Thomas Kuhn’s work were couched more in terms of mass psychology than of the unconscious, rational or un-rational. There is perhaps a further reason for the deep suspicion that surrounds the idea of the unconscious in science, one not dispelled by converting it into an Ørstedian rational unconscious, rather than a Freudian irrational one. For Ørsted, as for Helmholtz, the workings of the rational unconscious were intimately linked to pleasure, albeit a pleasure whose sources were ultimately inexplicable. If there is still a tabu so deep as to be unspoken in the history and philosophy of science, it is about pleasure, most especially the pleasures of the inexplicable. Max Planck Institute for the History of Science, Berlin 43

44

Jacques Hadamard, An Essay on the Psychology of Invention in the Mathematical Field [1945] (New York: Dover, 1954), pp. 12–14; 21–42. Arthur Koestler, The Sleepwalkers [1959] (New York: Grosset and Dunlap, 1963), pp. 334–336; 520–524.

ROMANTICISM AND RESISTANCE: HUMBOLDT AND “GERMAN” NATURAL PHILOSOPHY IN NAPOLEONIC FRANCE MICHAEL DETTELBACH

Celui qui ne s’occupe pas de l’univers, en Allemagne, n’a vraiment rien à faire. —Mme. de Staël, De l’Allemagne, 1ère partie, ch. XVIII

Any attempt to deal with the relationship of French science to German science in the Romantic period has to come to grips with the increasingly ethno-national self-consciousness that comes to pervade intellectual and cultural work in the Revolutionary and Napoleonic decades. Artists, poets, and critics measured their work against national origins. They cast themselves in a history of Europe defined by the mixture, conflict, and succession of peoples, each animated by a particular character or spirit, reciprocally determining and determined by physical environment, political institutions, religion, and art.1 So too did natural philosophers.2 Neglecting this dimension of cultural activity in the period runs the risk of attributing the prominent role of national categories in philosophical discourse to actually existing national differences and missing their internal, discursive function in the definition of nationhood in this critical period. In particular, we risk losing the constitutive role that scientific activity and the increasingly explicit discussions of “scientific method” played in the construction of national identities in the period. As Frederick Gregory notes in his essay in this volume, Ørsted was preoccupied from the 1800s on with concerns about defining an authoritative approach to knowing Nature, concerns which we tend to associate more with the 1830s and 1840s and the British context. Implicit in Gregory’s account of Ørsted’s debate with the Danish theologian and popular educator Frederik Grundtvig, and explicit in Andrew Jackson’s detailed exposition of the lateEnlightenment Danish church controversies that motivated Ørsted’s first ventures into natural philosophy, is that Ørsted’s “early” concern with defining a “scientific method” reflects a much broader contemporary debate about the place of church

1 2

Peter Hanns Reill, The German Enlightenment and the Rise of Historicism (Berkeley, 1975). Dietrich von Engelhardt, Historisches Bewußtsein in der Naturwissenschaft von der Aufklärung bis zum Positivismus (Frankfurt and Munich, 1979) in general makes this point but concentrates on the philosophies of history developed in Romantic physics, biology, and chemistry, and not on the historiographical discourse that formed the more general cultural context for their work.

247 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 247–258. © 2007 Springer.

248

M. DETTELBACH

and religion in the definition of nation and state. When we gloss over the national distinctions that permeate contemporary debates about proper philosophical “temper” and “method,” we miss the role of religion and religious debate and of a fundamentally theological interpretation of history. When our actors made distinctions between “French” and “German” science, they were defining nationhood in terms of attitudes toward nature, God, and the foundations of knowledge, the authority of personal experience and the credibility of personal witnessing. This paper argues for the importance of the religious question in Revolutionary and Napoleonic Europe—what should be the relationship between national church and national state? how is religious liberty best safeguarded?—in understanding the reception of German Romantic science in early 19th-century France. The question was salient in France, where the Revolution was often fought over clerical monopolies and religious freedom, and everywhere France extended its rule, not least into largely Protestant Swiss cantons and Germany. The question was equally salient in those territories forced by the political and military pressures of French conquest to define the basis of their own political identities—that is, all of Europe, its southeastern Ottoman fringes, and its colonies. When in 1812 Grundtvig attacked France’s emperor and its mathematical approach to Nature as anti-Christian, he was in part attempting to define a post-Napoleonic Denmark. The same concern for national identity permeated the relationship between French and German science during the Revolutionary decades, when Mme. de Staël almost single-handedly invented “Romanticism” as specifically German in protest against Napoleonic cultural and religious policies.3 In a series of works running through De la littérature et ses rapports avec les institutions (1803) and culminating in De l’Allemagne (written in exile between 1799 and 1807, though not published until 1810), de Staël used a carefully crafted image of German institutions and letters as a foil for “French” chauvinism, dogmatism, and absolutism, philosophical and political, and labelled it “Romanticism.” “The French and the Germans lie at two ends of the moral spectrum,” she opened De l’Allemagne. “Where the first consider external objects as the cause of all ideas, the second consider ideas the cause of all impressions.”4 Within this polarity, de Staël embraced poetry, literature, philosophy, sculpture, music, painting, and the institutions that governed them, universities and academies. She gave French audiences pioneering interpretations of Goethe, Schiller, and Kant, carefully tailored (to put it charitably) to contrast French moral slavishness and German moral freedom. In early 19th-century Paris, “German science” meant above all the science of Alexander von Humboldt. De Staël never directly addressed les sciences exactes et naturelles and only treated natural philosophy in the context of Kantian idealism, but there is no doubt that Humboldt’s work both contributed to and was interpreted through the same polemics that determined de Staël’s creation of Romanticism. 3

4

John Claiborne Isbell, The Birth of European Romanticism (Cambridge: Cambridge University Press, 1994). Isbell demonstrates that de Staël’s work systematically distorted her literary and philosophical sources to create “Romanticism” as a unified national counterpoise to Napoleonic France, and that her propagandistic synthesis defined Romantic revolt throughout the European world, including America, after the Restoration. De l’Allemagne 1: 19.

ROMANTICISM AND RESISTANCE

249

Through his elder brother Wilhelm von Humboldt, who corresponded intimately and at length with Germaine de Staël during her exile and directly shaped her interpretations of German poetry and philosophy, Alexander informed the construction of the distinction between national minds and tempers, even while working easily across them.5 Humboldt spent two decades in Paris between 1804 and 1827 and after returning to Berlin, made repeated extended stays well into the 1850s. On these sojourns he functioned as scientific and diplomatic mediator, conveying for instance the constitution of the Institut National to the Berlin Akademie in 1806, and the Observatoire’s metric standards to Berlin for the government of a considerably aggrandized Prussia in 1816.6 He had many of his works published in both German and French (one reason for insisting on publishing through the House of Cotta in Strasbourg), though save for the Essay on the geography of plants he never translated his own work; he wrote only in one language or the other. Examining the production and reception of Humboldt’s work in France in the first decades of the century offers an ideal way to explore the meaning of “German” natural philosophy in France in the period. When Humboldt returned triumphantly to Paris in the autumn of 1804, after five years in the Americas, the all-embracing character of his undertaking was much in evidence and the subject of much public discussion. He returned with a truly encyclopedic enterprise, encompassing physics, meteorology, and natural history, as well as archaeology, philology, and political economy; records of barometer readings and satellite immersion times, crates of dried plants and preserved monkey skins, as well as drawings of Aztec and Inca monuments, topographical surveys, and statistics on population and silver exports. Elected a foreign correspondent (correspondant étranger) of the Institut National just before landing in Bordeaux, he appeared before the assembly two or three times a week over the course of the 1804–1805 semester, exhibiting maps and plants and reading mémoires on subjects as diverse as Andean geology, crocodile physiology, and Aztec mythology.7 “I’m creating something of a frenzy [in the Institut],” Humboldt wrote to his brother Wilhelm in October, 1804. “You can see the gears spinning furiously in their heads, since I have often held forth on astronomical, chemical, botanical and astrological matters in the greatest detail, all in a single session.”8 Humboldt’s enterprise involved the collaboration of the most important scientists and scientific institutions of the capital: Berthollet, Laplace, Delambre, Cuvier, Biot, Gay-Lussac, Latreille, Riche de Prony, Langlès, Silvestre de Sacy,

5

6

7

8

For Wilhelm von Humboldt’s contribution, see Isbell, op cit. In the late 1790s, Wilhelm von Humboldt was himself busy composing a “panorama of the 18th Century” on the carpentry of national tempers, and spent much of his life articulating the privileged relationship between German and Greek language and literature. Archiv der Berlin-Brandenburgischen Akademie der Wissenschaften; Preussisches Geheimes Staatsarchiv (Dahlem), I.Rep.120. Abt.A, Fach IX [Mass und Gewicht], 1, Nr. 2, vols. 1 and 2, ff. 20–66: Anschaffung französischer und westphälischer Normalmaße und -gewichte. Humboldt was elected correspondant to the physical section on 16 pluviose XII. Procès-verbaux des séances de ‘Académie des sciences (La Hendaye, 1921), vol. 3, pp. 62, 171, 174. Paris, 14 October, 1804. Wilhelm und Caroline von Humboldt in ihren Briefen, edited by Anna von Sydow (Berlin, 1907), ii: 265–266.

250

M. DETTELBACH

Letronne; the Jardin des Plantes, the Bureau des Longitudes, the École polytechnique, the Louvre, the Société d’Arcueil. The Voyage de Alexandre de Humboldt et Aimé Bonpland aux régions équinoxiales du Nouveau Continent—eventually over 30 grand-quarto and grand-folio volumes, the largest publishing venture of its time, outdoing even the imperial Description de l’Egypte—sustained an army of young French physicists, mathematicians, draughtsmen, engravers, cartographers, while exhausting the resources of three publishing consortia. In 1809, the chemist and former minister Chaptal only dissuaded the Emperor Napoleon from ordering Humboldt deported as a Prussian spy by warning that, should the traveller be forced to leave Paris, science in the capital would come to a halt.9 Berthollet might very well have actually said what Humboldt claimed he did in the above-cited report to his brother: Cet homme réunit toute une Académie en lui.10 This pillar of the Parisian savant, salon, and publishing world, was decidedly Germanic. For all his vaunted cosmopolitanism, Alexander von Humboldt was identified and identified himself as particularly German, precisely because of his cosmic and cosmopolitan ambitions. Though written in French and composed in Paris, his dynamic physics pointed unmistakeably over the Rhine, to the methodological strictures of Kant and the nature-philosophy of Schelling. Humboldt justified his encyclopedic enterprise by describing the universe as a “general equilibrium of forces,” a dynamic Zusammenwirken der Kräfte: The general equilibrium which reigns amidst disturbances and apparent turmoil, is the result of an infinity of mechanical forces and chemical attractions balancing each other out. Even if each series of facts must be considered separately to identify a particular law, the study of nature, which is the greatest problem of general physics, requires the reunion of all the forms of knowledge which deal with the modifications of matter.11

Truly to understand Nature, that is, for physics to obtain its goal, “every physical phenomenon and every product of nature must be studied individually, then contemplated together with every other.” In his first plenary address to the Institut, an outline of a geography of plants, Humboldt exhorted his colleagues to raise themselves to “general views,” to realize a truly “philosophical” physics by bringing all the sciences together and discovering the “great laws of nature.” No fact or phenomenon could be viewed in isolation, he told the readers of his Tableau physique, because they were all connected dynamically. Nature was this nexus of forces.12 When in that inaugural lecture Humboldt called on his Parisian colleagues to pursue a “truly philosophical knowledge of nature” that would reveal the laws of the distribution and development of nature and civilization, they heard his German accent. Humboldt himself construed his science in national terms. Even as he prepared to leave Paris to find his intellectual fortune in the tropics in June, 1798, Humboldt 9 10

11 12

J. A. Chaptal, Mes souvenirs sur Napoléon (Paris: Libraire Plon, 1893), pp. 382–383. Paris, 14 October, 1804. Wilhelm und Caroline von Humboldt in ihren Briefen, ed. Anna von Sydow (Berlin, 1907), vol. 2, pp. 265–266. Chaptal recalled telling Bonaparte the same thing: “M. de Humboldt possède toutes les sciences, et lorsqu’il voyage, c’est toute l’Académie des sciences qui marche…” Op. cit. p. 383. Humboldt, Essai sur la géographie des plantes, pp. 41–42. Ibid. pp. 4, 44–45. The Essai was read before the Institut in January 1805.

ROMANTICISM AND RESISTANCE

251

contrasted the “German” approach to nature the “French.” He recognized that the French lacked a German sensibility to nature’s dynamic unity, and bought themselves religious and political order at the price of true insight into nature. He wrote his brother, who had solicited a characterization of French natural science for his tableau of the 18th century: In all the natural sciences, the French have a feeling only for mechanical and atomistic theories, and never for actual force and effect […] This keeps them safe from spiritualistic deviancies, but it also prevents them from getting to the root and from attaining a complete, natural view of things.13

Conversely, Humboldt conceived of his dynamic and unified view of nature “as animated by a single life,” his Ansichten der Natur, as a specifically German work. He composed the texts with explicit attention to the assonances and rhythms of the German language, and dedicated the book to his countrymen during the French occupation, deprived of any self-determining public life, so as to remind them of the great laws of nature and their own moral freedom. Although he insisted that Cotta publish German and French editions of his Essai sur la Géographies des plantes simultaneously, Humboldt carefully tailored his German translation of the original French to address fundamental “national” differences in the sciences. In particular, for the German edition, Humboldt expanded his preface to accommodate a lengthy acknowledgement of Schelling’s Ideen zu einer Philosophie der Natur (1798), which Cotta had forwarded to him the moment Humboldt returned to Paris in 1804. Humboldt deemed Schelling’s subsumption of phenomena under conceptual hierarchies “a physical portrait of a completely different, altogether higher kind” than his own, and openly encouraged Schelling’s attempts to find the dynamic unity of phenomena which a coarser physics, with less respect for the demands of human reason, considered so many separate substances. Who therefore can take happier and more heartfelt interest in a system which, undermining atomism and far from that way of thinking that reduces every difference in matter to mere differences in size and density… promises to shed considerable light on the phenomena of life, heat, magnetism and electricity, so inaccessible to science up until now?14

In the French edition, studying plants from a higher, philosophical point of view, in their geographical relationships, promised to advance physique générale, which referred to an encyclopedic program of mathematical construction of physical laws through precise measurement, and had deterministic and even atheistic impulses.15 To German readers of Ideen zu einer Geographie der Pflanzen, however, the same higher, geographical study promised to yield insights into a development “history of our planet” (Geschichte unserer Planet). French readers expected 13 14

15

Wilhelm von Humboldt, Gesammelte Schriften, XIV: 585–586. Alexander von Humboldt, Ideen zu einer Geographie der Pflanzen nebst einem Naturgemälde der Tropenländer. In Schriften zur Geographie der Pflanzen, edited by Hanno Beck (Darmstadt: Wissenschaftliche Buchgesellschaft, 1989), 44–45. See also Humboldt’s enthusiastic correspondence with Schelling from 1805 in Briefe der deutschen Romantiker, edited by Willi A. Koch (Leipzig: Dietrich Verlagsbuchhandlung, 1938), pp. 201–204. Frankel, Eugene. “J.B. Biot and the Mathematization of Experimental Physics in Napoleonic France.” Historical Studies in the Physical Sciences, 8 (1977), pp. 33–72.

252

M. DETTELBACH

Humboldt’s plant geography to lead to increasingly precise mathematical relationships between physical variables, in the manner of Biot and Laplace; German readers expected history, not theory: an evolutionary vision of the history of the earth, both nature and Man. Plant geography would enable the German reader to understand the particulars of natural and civil history as the working-out of a single, dynamic process. As the Emperor’s 1809 threat to deport the Prussian savant suggests, the imperial capital was not an entirely congenial home for Humboldt’s cosmic ambition. Alarmed Parisian critics, even among his collaborators, accused Humboldt of dilletantism, of forsaking scientific rigor for the undisciplined breadth of the amateur. Berthollet panned Humboldt’s chemical measurements as figments of his imagination; Laplace questioned his measurements of terrestrial refraction and his claim to have discovered a sophisticated astronomy among Aztec ruins; Ramond de Carbonnières impugned the precision of his barometric measurements.16 This is not to say that the leading French savants were hostile to Humboldt. On the contrary, Biot, Berthollet, Laplace and many others enjoyed and profited from the wealth of carefully collected material Humboldt brought back to Paris, and they associated closely with Humboldt in the Society of Arcueil and at the Institut. But they took up a bemused and sometimes hostile position toward the cosmic embrace of Humboldt’s scientific work, and his public persona as himself a force of nature. On dit souvent en société que je m’occupe de trop de choses à la fois, Humboldt complained to the Genevan naturalist Marc-Auguste Pictet in 1806. Humboldt vigorously rejected the charge: Je réponds: peut-on défendre à l’homme d’avoir le désir de savoir, d’embrasser tout ce qui l’environne? […] Et que l’on examine si, dans les petit* essays que j’ai faits des différents branches [des sciences], je n’ai pas eu la constance de poursuivre le même objet. Et pour avoir des vues générales, pour concevoir le liaison de tous les phénomènes, liaison que nous nommons Nature, il faut d’abord connaître les parties, et puis les réunir organiquement sous un même point de vue.17

It was a view of science which few of his Parisian colleagues shared, even as they marvelled at his facility with their instruments. In their skepticism, Humboldt’s critics among the Parisian savants reflected a more general, positive hostility to German philosophical and religious “excess.” 16

17

Berthollet: “Observations sur l’action que le sulfate de fer exerce sur le gaz nitreux,” Annales de chimie, 39 (An 9 [1801]): 3–17; Laplace to Humboldt, Paris, 8 May 1806, UB Leipzig; Ramond regretted that Humboldt had only measured his barometric altitudes to 1/10th of a line, or +/− 3 meters. Journal de Physique 60 (1805): 280 ff. Ramond’s continuing efforts to correct Laplace’s formula for calculating elevations from barometric frustrated Humboldt no end: “Ramond a lu de nouveaux mémoires sur ces éternels Baromètres. Il y a des méchans qui noyent au coups d’épaules [?]. Il trouve des millimètres de hauteur et finira par mesurer les conscrits au moyen du Barometre, ce qui rendra la méthode tres recommendable sans doute.” (to Pictet, Paris, 30 December [1812?], Fondation Rilliet, Geneva, no. 26) To Humboldt, Ramond’s Laplacian insistence on mathematical completeness and perfection was a prime example of the French penchant for atomism and mathematical realism, and their insensibility to a dynamic view of nature. Humboldt to Pictet, 1806. Stapfer reported to Usteri (14 October 1804) on the wide variety of results Humboldt was promising soon after his return and the skepticism they were meeting among French scientists. “Allein ich sehe, dass mehrere hiesige Gelehrte in die Genauigkeit und Zuverlässigket dieser Angaben und Resultate kein völliges Zutrauen setzen,” citing specifically Humboldt’s eudiometric results. Stapfer’s Briefwechsel (Basel, 1891), i: 174.

ROMANTICISM AND RESISTANCE

253

Paradoxically, as the reach of Paris extended over Europe, even to the Vatican museums and the blessing of the Pope Pius VII, spirits contracted. François Guizot, a young journalist at the time, recalled the constriction of public life during the Empire in his Mémoires: “L’enivrement de 1789 avait bien complètement disparu. […] La sécheresse, la froideur, l’isolement des sentiments et des intérêts personels, c’est le train et l’ennui ordinaires; la France, lasse d’erreurs et d’excès étranges, avide d’ordre et de bon sens commun, retombait dans cette ornière.”18 Independent minds (“educated and somewhat sensitive to the fate of human dignity,” Guizot called them) retreated within the bounds of private interest and the formalities of state service. To its opponents, official French culture had sunk into a bland neoclassicism and a militant chauvinism in letters and criticism, and a narrow utilitarianism in the sciences. “In France, the ignorance of values and customs among other peoples is taken to the outer limits of ridiculousness,” reported one German observer (the Dresden theatrical impresario J. F. Reichardt) living in Paris under the Consulate.19 German philosophy, art, religion, and literature were singled out for neglect, and their introduction actively thwarted by official Napoleonic culture, precisely because Germany allegedly purveyed a universalism contrary to Imperial ideas of French hegemony. “Some people accuse me of having opinions they consider ‘Germanic,’ ” wrote Guizot editorially in Le Publiciste in 1809. Of all European nations, the Germans have left open the freest channels for all that comes from without; convinced that the works of the human spirit wherever they flower are the patrimony of all mankind, they are constantly concerned to profit from this legacy. … [They possess] a general tendency that makes the Germans a nation truly foreign to all literary egoism, truly cosmopolitan in its works.20

Between 1802 and 1815, there were numerous attempts to introduce Kant’s philosophy, Goethe’s and Schiller’s poetry and criticism, Heyne’s philology and biblical criticism to French readers, mostly in vain.21 18

19

20

21

Guizot, Mémoires pour servir à l’histoire de mon temps, 8 vols. (Paris: Michel Lévy, 1858–1867), vol. 1, p. 7. Guizot also made ends meet in Paris as tutor to Stapfer’s children, 1807–1811. “L’ignorance des moeurs et des habitudes de la vie chez les autres peuples est poussée en France jusqu’aux dernières limites du ridicule,” J. F. Reichardt, Un hiver à Paris sous le Consulat, 1802–1803 (Paris, 1896), p. 87. “Quelques personnes me reprochent des opinions qu’elles appellent germaniques. Les Allemands sont, de toutes les nations europénnes, celle qui a laissé les avenues les plus libres pour tout ce qui lui venoit du dehors; convaincus que les travaux de l’esprit humain, en quelque lieu qu’ils aient pris naissance, sont le patrimoine de tous les hommes, ils sont toujours empressés de profiter de cet héritage précieux… [Il y a] une disposition générale qui fait de la nation allemande une nation vraiment étrangère à tout égoisme littéraire, vraiment cosmopolite dans ses travaux.” Le Publiciste, 5 August and 29 August 1809; quoted in Wolfgang Leiner, Das Deutschlandbild in der französischen Literatur, p. 266. In general, Roland Mortier, Les “Archives littéraires de l’Europe” (1804–1808) et le cosmopolitisme littéraire sous le premier Empire (Brussels: Palais des Académies, 1957). For instance, in 1804 the Institut created a commission (including Humboldt) to establish a Bibliothèque germanique on the model of the Bibliothèque britannique edited by the Genevan naturalist (and former tribune under the Consulate) Marc-Auguste Pictet, but the interior minister objected to every editorial board put forward, and the project never moved ahead. (Procès-verbaux des séances de l’Académie, iii: 164, 173–174) Stapfer’s and Villers’ attempt to publish a “Mélanges de littérature étrangère” likewise came to nothing, “eben weil wir nur reine Humanitätszwecke im Auge haben.” Stapfer Briefwechsel, i: 173.

254

M. DETTELBACH

The literary cosmopolitanism which was perceived as characteristically German in Napoleonic Paris did not simply offend French imperial chauvinism. It was politically threatening, because it attributed to the individual mind the power to embrace the entire universe, to move from local appearances to universal forces and laws. God was within the purview of every man’s reason and experience. According to a German philologist residing in Paris in 1805, the Emperor Napoleon believed “that the Germans do not do anything, not even chemistry and physics, without mixing in politics, liberty, and revolution.”22 Universalism was central to the idea of Germany and the German at the turn of the 19th century. This emerged especially clearly in debates over what to do about Protestant Germany’s famously fractious universities, and how to structure a new French university system. Celui qui ne s’occupe pas de l’univers, en Allemagne, n’a vraiment rien à faire, wrote Mme de Staël in the chapter on German universities in De l’Allemagne. French reformers debating the form and structure of an imperial university and critical of the organization of higher education in France specifically looked to German universities as producers of universal men, and contrasted them with French schools designed to supply state functionaries. Charles de Villers, de Staël’s fellow exile, vigorously defended the universities of northern Germany to their new sovereign, Jerôme Bonaparte, king of Westphalia, as the Emperor Napoleon ordered their integration into the imperial university system. To understand and evaluate a German university, one must completely disabuse oneself of all the stereotypes of ordinary schools, of monastic order, and of that collegiate discipline imposed on children. Here these are men talking with men.

In the universities of protestant Germany, wrote Villers to Jérôme: all the sciences bear on one another and are linked to one another by tight chains that cannot be broken without prejudice. This is what above all makes this form of university, which comprehends the entire cycle of learning seem preferable to that which consists of special schools or separate faculties, which exists in France. [In the German university] it is difficult for one to be purely a lawyer, a doctor, or a man of letters. One who has received only a narrow education strictly in a single science will always lack the general views, the broader knowledge which links his science to all the rest of human knowledge, which completes, elevates, or ennobles it.23

22

23

“que les allemands ne s’occupent de rien, pas même de chimie et de physique, sans y mêler la politique, la liberté et la révolution.” Hase to Böttiger, 11 March 1805. Hase was diagnosing the failure of a planned monthly digest of German literature and philosophy. See Ludwig Geiger, “Eine deutsche Zeitschrift in Frankreich (1805),” Zeitschrift für vergleichende Literaturgeschichte, 10 (1896): 350–352, 495–495; Joseph Texte, “Les origines de l’influence allemande dans la littérature française du XIXe siècle,” Revue d’Histoire littéraire de la France, 5e année (1898), pp. 1–53. Charles Villers, Coup d’oeil sur les universités et le mode d’instruction publique de l’Allemagne protestante (Cassel, 1808), quoted in Texte, pp. 20–22. “Il faut absolument se défaire, pour concevoir et juger un tel institut [une université allemande], de toute arrière-pensée d’école ordinaire, de régularité monastique, et de cette discipline de collège qu’on impose à l’enfance. Ce sont ici des hommes qui parlent à des hommes. Toutes les sciences s’appuient mutuellement et se tiennent par une chaine étroite qui ne peut se rompre sans préjudice. C’est par où surtout la forme des universités qui embrassent tout le cycle de l’enseignement nous parait préférables à celui des écoles spéciales ou des facultés séparées qui en tiennent lieu en France. Il est difficile qu’on soit tout purement jurisconsulte, ou médecin, ou lettré. Il manquera toujours à celui qui n’aura reçu qu’un enseignement strict et exclusif dans une science, les vues générales, les connaissances accessoires qui lient sa science à tout le reste du savoir humain, qui le complètent, le relèvent ou l’ennoblissent.

ROMANTICISM AND RESISTANCE

255

Like Humboldt’ s “view of nature” with its universal scope and attention, German universities aimed to cultivate vues generales and the moral and intellectual independence they afforded. In protestant Germany, universities created “whole men,” capable of making moral decisions and thinking for themselves, claimed Villers. At the same time, in France, observers saw an educational system ossifying into ever-narrower channels designed to produce men of narrow skills. When in 1811 Napoleon granted preference in state service to graduates of military colleges over graduates of the École Polytechnique, and simultaneously eliminated the last elective elements in the poytechnical curriculum, it only confirmed what many had long known: “The principle that the sciences have no purpose other than to produce gunpowder, sugar, indigo, cotton etc. more cheaply is becoming ever more explicit.” (Das Axiom, dass die Wissenschaften keinen anderen Zweck haben, als Pulver, Zucker, Indigo, Baumwolle u.s.w. auf wohlfeilerem Wege zu erhalten, spricht sich immer deutlicher aus.)24 The moral freedom which Villers and others saw being cultivated in the German university is what Humboldt advertised as the ultimate payoff of studying nature. This is most explicit at the end of his introductory lecture on plant geography, where he contrasts the savage in the tropics, surrounded by all of nature’s forms but unaware of their lawfulness, and the cultivated individual isolated on a desert coast, but able through the gifts of civilization to recreate the living universe in his mind. In Europe, the individual isolated on an arid coast can delight in the aspect of distant regions, in his mind. If his soul is sensible to works of art, of his cultivated spirit is sufficiently extended to raise itself to the grand conceptions of general physics, from the depths of his solitude, without moving from his place, he takes for himself all that which the intrepid naturalist has discovered in travelling the oceans and the skies, in delving into subterranean caves, or in mounting icy summits. This is how, without a doubt, enlightenment and civilization [literature and the imitative arts] most influence our individual happiness. […] It is through these researches that we open the way to intellectual delight, to a moral liberty which fortifies us against the blows of destiny and which no exterior power can undermine.25

The figure of Humboldt and the Humboldtian naturalist, this civilized nomad or sociable solitary projected by Humboldt himself and reflected by his audience in Paris, had clear political resonances. Humboldt’s “views of nature” were supposed to inspire the reader with the same stoic moral authority he displayed in his mastery of nature and history, and which Villers saw as threatening to, and threatened by, the French imperial state. Leibniz was emblematic of German philosophy and its claims for the selfsufficiency and in-principle universality of the individual mind. “That which once and for all established [Leibniz’s] glory,” Mme. de Staël summed up in her introductory chapter on German philosophy, “is that he knew how, in Germany, to maintain the philosophy of moral liberty against that of sensualist fatalism.”26 24 25 26

Stapfer Briefwechsel, ii: 43. Humboldt, Essai sur la geographie des plantes, pp. 34–35. DA 3: 5. De Staël shared with the Institut’s Class of Moral Sciences this interpretation of Leibniz as well as Kant principally as moralists. See Mortier, op cit. pp. 154–173.

256

M. DETTELBACH

When the Zurich statesman Philippe Albert Stäpfer, anonymously reviewing the first volume of Humboldt’s Relation historique du voyage aux régions équinoxiales du Nouveau Continent (1814) in the official Moniteur universel, yielded to enthusiasm for Humboldt’s grandes vues and praised the Prussian as a modern “Leibnitz,” Humboldt rushed to silence him: “With such a noble and simple soul, you have forgotten for a moment that it is not permitted to make comparisons with Leibnitz. It is the time of terror against foreigners.”27 In Napoleonic Paris, Humboldt’s particularly “German” universalism was intellectually and politically suspect. The last thing the emperor wanted was a new “Leibniz,” encouraging cosmpolitanism, and insisting on the rational comprehensibility of the entire universe, on the ability of the individual mind to grasp the cosmos. This Leibnizian ability to see a world in a grain of sand is just what some chaffing Parisian savants and salonnieres appreciated about the German naturalist in their midst. In an 1808 review of the French translation of Humboldt’s Ansichten der Natur (Tableaux de la nature, 1808), Georges Cuvier contrasted the naturaliste voyageur and the naturaliste sédentaire, the field naturalist and the cabinet naturalist: the one able to witness nature’s “grand scenes” and seize objects “in the very places they were placed by nature, in their true relation with their environment, and in all the fullness of life and action”; the other able “through reflection, observation, and erudition to illuminate an object with all the light the present state of knowledge can provide.” Humboldt united these two types dans un seul homme. His ability at once to convey the grand objects of nature in the greatest detail and vividness, and yet to bring the full weight of learning and human experience to bear in order to present them as the product of general laws, made him the model student of nature. “When he presents his reader with the grand views of nature, he seems always to have reflected; when he brings together data, recalls and weighs opinions, he seems never to have left the library; when he traces the sketch of his grand results, he seems to have given himself over ceaselessly to meditation.” Humboldt’s ability to experience Nature vividly without falling into disjointed sensualism, to see “Nature as animated by a single life,” contrasted starkly with the inability to rise above the particular and immediate which Cuvier took to be symptomatic of modern life, and especially French life under the Empire.28 Humboldt’s dynamic “views of nature” elicited views of history from Cuvier, the 27

28

Humboldt to Philippe Albert Stapfer, “ce mardi” [24 January 1815], Wellcome Institute for the History of Medicine, MS#40. (mistakenly dated “1809”) The review is in Moniteur universelle, 22 janvier 1815. Stapfer was the Helvetic representative to Bonaparte’s court. Rodolphe Luginbühl, “Alexander de Humboldt et Philippe Albert Stapfer,” Denkschrift der historischen und antiquarischen Gesellschaft zu Basel (Basel: Schweigerhauserische Buchdrückerei, 1891) recites the circ*mstances surrounding the review and partially quotes Humboldt’s response, pp. 162–168. Cuvier’s Humboldt closely resembles the heroes and heroines of de Staël’s novels—all nomads or barbarians of the drawing room—as well as Cuvier’s own self-image (see n. 31 below). Isbell, op cit. pp. 50–53. The two touchstones for this double-movement of sensation and reflection, turning outward and turning inward, in the German literature are Schiller’s Uber naïve und sentimentalische Dichtung (1796) and the second part of Kant’s Kritik der Urtheilskraft (1791), where Kant explains true sublimity not as terror before nature’s immensity, but as a final realization of the human soul’s superiority to nature. Stapfer’s praise of Humboldt as “Leibniz and Cook dans un seul homme” (in a mistaken eulogy of 1811) reflects a similar appreciation of Humboldt as a symbol of moral liberty.

ROMANTICISM AND RESISTANCE

257

struggle of human societies against nature and against one another. Specifically, Humboldt’s views prompted Cuvier to reflect on how constricted and mechanical the effort of civilization had made life in Europe. The division of labor rivets every man to his office, his workshop, or his plot of land; nothing varies each man’s work or stimulates his reflection. Thence a different cast of mind: a complete dependence of men on one another, and values which resemble those of the ancient Greeks as little as our cabaret singers resemble Homer or our philosophers Socrates and Plato. Thence also an entirely different mode of policing, in order to maintain and direct this great manufactory, in which no one retains his individuality and no one may exercise his own will.29

This was the passage which, I think, prevented the review’s publication; it remains a manuscript in the Fonds Cuvier. Cuvier was closely tied to the Napoleonic regime by his positions as professor of comparative anatomy at the Museum and, more importantly, as inspector-general and member of the directing council of the new Imperial University. But he also kept his intellectual distance from the Napoleonic state, not least by virtue of his Protestant, Montbéliard heritage.30 (It was Cuvier who, as a permanent secretary of the Institut, prompted the 1802 prize question on “the influence of the Lutheran Reformation on the political situation of the different states of Europe and on the progress of knowledge?”31) He regarded Humboldt’s encyclopedic physics as a model of intellectual liberty in an increasingly centralized and newly re-Catholicized regime.32 Like Humboldt and Ørsted, Cuvier manifested an early and deep preoccupation with issues of method in natural science, springing from the same commitment to religious liberty. Humboldt repeated this assertion of intellectual liberty to Napoleon himself in presenting an inscribed copy of the Tableaux de la nature to the Emperor in February, 1808: “Imbued with sentiments of gratitude and admiration which are

29

30

31

32

Bibliothèque de l’Institut de France, Fonds Cuvier, MS 3159, 2r–3v, 6v. La division du travail attache chacun à son bureau, à son atelier, où à son glèbe; rien n’y varie l’occupation des individus et n’excite pas la reflexion; de là une autre tournure dans les esprits: une dependance absolue des hommes entre eux et des moeurs qui ressemblent aussi peu à celles des grecs que nos chanteuses publics diffèrent a l’homère our nos professeurs de philosophie à Socrate et Platon; de là aussi une toute autre police pour maintenir et diriger cette grande fabrique ou personne ne garde son individualité et ne doit avoir jamais une impulsion propre. See Outram (1984), pp. 143–147, for Cuvier’s religious opinions, so far as they are known, and a brief summary of his role in official French protestantism as member of the nominating committee of the Protestant Consistory of Paris (under the Empire) and Grand-Master of the Faculties of Protestant Theology in the Université de France and Director of non-Catholic Religions under the Restoration. On the occasion of the Concordat of Bologna, Cuvier wrote to Charles de Villers, whose Essai sur l’esprit et l’influence de la Réformation de Luther (Paris, Metz, 1804) won the prize: “What say your Protestants and above all your Kantians about all the fine things we’re up to here? Here are our materialists [ideologues] who, wanting nothing to do with noumena and pure understanding, are now obliged to swallow transubstantiation with all its charms; besides, they say that a god made of bread suits them as well as any other, it’s all matter.” (18 floreal an X, quoted in Texte, p. 23). Outram (1984) describes the Romantic persona Cuvier crafted in the “Preliminary Discourse” of Recherches sur les ossem*nts fossils (1812), the solitary explorer of space and time, attaining views over natural and civil history. This persona is close cousins with the Humboldt of Ansichten der Natur and with the “solitary man” he conjures at the end of the Essai sur la géographie des plantes, who, though “isolated on an arid coast,” can nonetheless survey all of nature through the tools of civilization and science.

258

M. DETTELBACH

inspired by the generous interest and the protection which Your Majesty deigns to grant to the sciences that I study, I shall strive unceasingly for a life to which is attached not only the glory of the name of France, but the progress of the civilization of mankind.”33 The reports on the progress of the sciences which Cuvier and Joseph Delambre addressed to the Conseil d’État, as the permanent secretaries of the Institut’s Physical Section, betray a similar concern to assert the superiority of the sciences to the interests of the state. This was just five days after Humboldt’s Tableaux de la nature had landed on Napoleon’s desk. Though full of praise for the Emperor’s generous support and heavily stressing the accomplishments of French scientists, Cuvier concluded his report with a covert admonition: “an ordinary prince,” Cuvier told Napoleon, would have allowed immediate utility to dictate the course of science, but the Emperor clearly recognized that the sciences had their own, internal ends.34 The glory and power of France were not worthy or adequate measures of true knowledge. The ramifications of Humboldt’s reception and self-presentation in Paris in the early decades of the 19th century are only selectively and briefly sketched here, but they sufficiently indicate what contemporaries believed to be at stake in “German” Romantic science. Animating the interaction between French and German science was the perennial “religious question,” put to Europe with renewed sharpness by the French Revolution. Behind the concern with scientific attitudes and casts of mind, and the tendency to define them along national lines, was a debate on intellectual and religious freedom. Science played proxy for religion in these polemics. Debates about “scientific method” and the definition of “science” and “scientist” were moral ones, and their growing explicitness and discursive independence after the Restoration and into the 1830s signal in many ways the enduring legacy of Romantic science rather than its extinction. Boston University

33

34

Humboldt to Bonaparte, 1 February 1808, Wellcome Institute for the History of Medicine, London. Quoted in Maurice Crosland, Society of Arcueil, pp. 44–45. Cuvier admired German “cosmopolitanisme,” where “on s’attachait a la justice universelle plus qu’aux interets parrticuliers d’un etat.” (quoted in Outram 1984, p. 69).

BETWEEN ENLIGHTENMENT AND ROMANTICISM: THE CASE OF DR. THOMAS BEDDOES TREVOR H. LEVERE

Dr. Thomas Beddoes is the main subject of this essay, but I shall work back to him via Davy, Berzelius, Coleridge, and Ørsted. Beddoes was a figure in the transition in England from the Enlightenment to Romanticism. His official biography,1 published by a dull doctor called Stock, suppresses almost everything that lent excitement to his life—his political activism, his conflicts with authority, his gift for friendship, his energetic internationalism, his scorn for the establishment, and the exuberant breadth of his intellectual, professional, and frequently radical connections. Besides labouring under suspicion from the Home Office, incurring the hostility of Joseph Banks, expressing enthusiasm for liberty, equality, and fraternity, and wishing for democracy in England, Beddoes was a conduit for European (and especially German) scientific, medical, philosophical, and literary culture. He was also, and to good effect, a patron of brilliant youth. He was physician, friend, co-agitator, and intellectual stimulus to Samuel Taylor Coleridge, and, as director of the Pneumatic Institution at Clifton, just outside Bristol in the west of England, provided young Humphry Davy with a high-level entry into his stellar chemical career. In 1812, more than a decade after leaving Beddoes’s Institution, Davy published his Elements of Chemical Philosophy, describing it optimistically but vainly as Part 1, Volume 1.2 That was all that ever appeared. Davy had sent a copy of this book to the Swedish chemist Jöns Jacob Berzelius, who undertook to send him comments. Berzelius, normally forthright in allotting praise or criticism, hesitated in this instance before writing. He observed that Davy had apparently studied chemistry more by experiment than by reading. No doubt that was the best way, but it meant that there were imperfections in those parts of the book where Davy’s laboratory experience did not reach. “Great minds”, he wrote, “often attach too little importance to details, and perhaps they are wrong to concentrate on anything other than wide and general views. Your philosophy is too far above my criticism; perhaps my head is too ill organized to be able

1 2

J. E. Stock, Memoirs of the Life of Thomas Beddoes M. D. (London and Bristol, 1811) H. Davy, Elements of Chemical Philosophy, Part 1, Vol. 1 (London: J. Johnson, 1812).

259 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 259–272. © 2007 Springer.

260

T. H. LEVERE to follow it. In hypothetical matters I limit myself to requiring of the author a precise and clear exposition of all the probabilities, and to wishing not to be prejudiced, by the expression of his opinions, in favour of any hypothesis. I feel a need to believe nothing without proof, and I want positive knowledge. I enjoy the study of probabilities, but wish to leave decisions about them to future research. I consider the ideas on light, caloric and electricity, as interesting fictions, which will dazzle less obstinate readers than me, and which will one day give way to the fictions of another century.”3

Davy’s book contained too much speculative theorizing for the hard-headed Swede, who, predictably, had already taken up the cudgels against Romanticism,4 and especially against its aggressive associate or partner-in-crime, the Naturphilosophie of Schelling. Romantic science and Romantic chemistry were not terms that Berzelius would have recognized—he would have recognized the words, but would have considered that the very idea of Romantic science was an oxymoron. In 1811 he had carried out a thorough demolition job on the electrochemical theories of the Hungarian chemist Jacob Joseph Winterl, who had set out his ideas in 1802– 1803.5 Winterl had argued for the elemental nature of water, and had presented positive and negative electricity as the acid and basic principles respectively. He had also come up with several new elements, including “andronia,” which other chemists failed to isolate. His results were not reproducible, and his theories were, in Berzelius’s view, worthless—a chaos that Berzelius found unsurprising, given its philosophical genesis in the work of Schelling.6 Berzelius’s objections were not shared by Hans Christian Ørsted, who in1806 had written a long paper on the series of acids and bases,7 in which he acknowledged what he termed Winterl’s beautiful demonstrations, and assumed, “at least de jure, that every chemist has read and pondered upon what he has said regarding this…. In general,” Ørsted noted,8 “in the course of this investigation we will frequently use what this profound researcher has done for this subject ….” Ørsted also made admiring reference to “a very beautiful series of experiments” on electrochemistry by Berzelius and Hisinger, in which they showed that acid appeared opposite the

3

4

5

6 7 8

Berzelius to Davy, late 1812 or early 1813, translated from Jac. Berzelius Bref, edited by H. G. Söderbaum, vol. 2, correspondence between Berzelius and Sir Humprhy Davy (1808–1825) (Uppsala: Almqvist and Wiksell, 1912), pp. 36–37. This episode is presented by Sven Lindroth, “Berzelius and His Time,” in Evan M. Melhado and Tore Frängsmyr, eds., Enlightenment Science in the Romantic Era: The chemistry of Berzelius and its cultural setting (Cambridge: Cambridge University Press, 1992), pp. 9–34 at p. 19. Prolusiones ad chemiam saeculi decimi noni (Budapest, 1800); Accessiones novae ad prolusionem suam primam et secundam (Budapest [1803]); Darstellung der vier Bestandtheile der anorganischen Natur: Eine Umarbeitung des ersten Theiles siener Prolusionen und Accessionen von dem Verfasser, translated by J. Schuster. See J. R. Partington, A History of Chemistry, vol. 3 (London: Macmillan, 1962), pp. 599–600. See also Anja Skaar Jacobsen, “Between Naturphilosophie and Tradition. Hans Christian Ørsted’s Dynamical Chemistry,” Ph.D. dissertation, Århus University, 2000, which has an important discussion of Winterl as well as of Ørsted, and also her paper “Spirit and Unity: Ørsted’s Fascination by Winterl’s Chemistry,” Centaurus 43 (2001), 184–218. Lindroth, op. cit. Ørsted, Journal für die Chemie und Physik, edited by A. F. Gehlen, 2 (1806), pp. 509–547. Ibid. in Selected Scientific Works of Hans Christian Ørsted, translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen (Princeton, NJ: Princeton University Press, 1998), pp. 227–242 at p. 229.

BETWEEN ENLIGHTENMENT AND ROMANTICISM

261

positive pole, and base opposite the negative.9 Berzelius,10 in spite of Ørsted’s enthusiasm for his work, was not overmuch impressed by Ørsted’s theorizing, or by the latter’s admiration for Winterl. Berzelius of course knew Ørsted. They first met in Copenhagen in 1807, began to correspond in 1810, and after a while became friends. Berzelius’s last foreign journey was to Copenhagen in 1847, where he stayed with Ørsted. Even had he not known Ørsted personally, he would have known of Ørsted’s writings, given his own omnivorous appetite for chemical publications, which made him the revered and feared arbiter of European chemistry through his weighty and authoritative annual reports on the progress of chemistry. His appetite for all chemical publications would in this instance have been reinforced by “the importance of Denmark to the intellectual life of Sweden at this period”, which Gunnar Eriksson has observed “has perhaps never been sufficiently emphasized.”11 When in 1812 Ørsted published his Ansicht der chemischen Naturgesetze, he again expressed his admiration for the profound discoveries of Berzelius, coupling him this time with Davy. Berzelius could not reciprocate; he had no time for Ørsted’s new nomenclature, nor for his speculations, and said so. Ørsted had offered his compliment to Berzelius in the same paragraph as a renewed endorsem*nt for Winterl, along with the statement that “Ritter can…be regarded as the creator of modern chemistry.”12 Berzelius was not impressed. Johan Wilhelm Ritter was the chemist closest to the Naturphilosophie of Schelling. Davy had castigated Ritter for his errors as a theorist, which “seem to be derived merely from his indulgence in the peculiar literary taste of his country, where the metaphysical dogmas of Kant which as far as I can learn are pseudo Platonism are preferred before the doctrines of Locke and Hartley, excellence and knowledge being rather sought for in the infant than in the adult state of the mind.”13 Clearly Davy had not read Kant, or at least not understood him, nor had he been keeping up with philosophical developments in Germany. He was however right to associate Ritter’s theories with German metaphysics, and with idealism. There was simply too much speculation. 1812 also, as we have seen, had Berzelius commenting on Davy’s Elements of Chemical Philosophy, complaining that Davy was overly speculative. This was wounding to Davy. Ritter was speculative, Winterl was speculative, Ørsted 9

10

11

12

13

Ibid. pp. 238–239, referring to Hisinger and Berzelius, Neues allg. Journ. d. Chemie, 1 (1803), pp. 115–119. For the relations between Ørsted and Berzelius, see Correspondance de H. C. Örsted avec divers savants, ed. M. C. Harding, 2 vols. (Copenhagen: H. Aschehoug, 1920), vol. 1, pp. 1–75. G. Eriksson, “Berzelius and the Atomic Theory,” in Enlightenment Science in the Romantic Era (1992), pp. 56–84 at p. 65. Ørsted, Ansicht der Chemischen Naturgesetze (Berlin, 1812), translated as The Chemical Laws of Nature in Selected Scientific Works of Hans Christian Ørsted, pp. 310–392 at p. 313. H. Davy, MS lecture, n.d. but about 1808, quoted in Levere, Affinity and Matter: Elements of Chemical Philosophy 1800–1865 (Oxford: Clarendon Press, 1971), p. 33. Ritter’s electrochemical researches have been reprinted: Johann Wilhelm Ritter, Entdeckungen zur Elektrochemie, Bioelektrochemie und Photochemie von Johann Wilhelm Ritter ; aus seinen Abhandlungen ausgewählt, eingeleitet und erlautert von Hermann Berg und Klaus Richter (Leipzig : Geest & Portig, 1986), Ostwalds Klassiker der exakten Wissenschaften, Bd. 271. See also Walter D. Wetzels, Johann Wilhelm Ritter, Physik im Wirkungsfeld der deutschen Romantik (Berlin, New York: de Gruyter, 1973), and Stuart Walker Strickland, “Circ*mscribing science: Johann Wilhelm Ritter and the physics of Sidereal man” Ph.D. Thesis, Harvard University, 1992.

262

T. H. LEVERE

was speculative—but Davy?! Davy considered that he himself belonged in more empirical and therefore more respectable scientific company. Samuel Taylor Coleridge, poet, critic, philosopher, self-styled library cormorant, former political activist, and one time admirer and friend of Davy,14 agreed with him on that score. Sometime between 1817 and 1820, Coleridge read and annotated Ørsted’s Ansicht.15 In annotating Ørsted, as he annotated and thereby, as he saw it, enriched almost all the books that he read, he would have provided some consolation to Davy, had he and Davy still been on close and friendly terms. The critical point for Coleridge came with Ørsted’s discussion of chlorine. Lavoisier, by applying analogy more rigorously than his definition of an element as the last product of analysis, had invented oxymuriatic acid. He believed that all acids were, like phosphoric acid, sulphuric acid, and carbonic acid, produced by the combination of a non-metal with oxygen, the acid-generating element. It followed that the acid obtained from sea-salt, our hydrochloric acid, must contain oxygen. Davy, having repeatedly and vigorously tried and failed to decompose the green gas that was one of the constituents of hydrochloric acid, concluded that he was dealing with an element, and he called it chlorine.16 Ørsted objected that the analogy between hydrochloric acid and nitric and sulphuric acids argued for the compound nature of chlorine. Coleridge sided vigorously with Davy: It is of highest importance in all departments of Knowledge to keep the Speculative distinct from the Empirical. As long as they run parallel, they are of the greatest service to each other: they never meet but to cut and cross. This is Ørsted’s fault—the rock of offence on which this Work strikes. Davy is necessarily right: for he follows the established Regula recta [right rule] of empirical Chemistry, viz. that all Bodies shall be considered as simple, till they have been shewn to be compound. On this Rule Chlorine, and Iodine claim the title of simple Bodies (Stoffen) with the same right as Oxygen or the Metals: while the Speculative Chemist sees a priori, that all alike must be composite.17

Coleridge believed that all chemical species, as the product or synthesis of immaterial polar powers, were essentially composite; but he also acknowledged that empirical chemistry should be bound by chemical rules of evidence. He had learned his chemistry largely from Davy, in conversation, by attending his lectures at the Royal Institution of Great Britain, and even, in the early days, by attending some of Davy’s pneumatic experiments as director of the chemical laboratory in Dr. Thomas Beddoes’s Pneumatic Institution. It was through Beddoes and his Institution that Coleridge and Davy first met one another. Beddoes, a former student of Dr. Joseph Black in Edinburgh, partner with James Watt in the design of pneumatic apparatus, and during the 1790s a vigorous critic of government policies towards France and an advocate of democracy, probably met Coleridge while attending a protest meeting against the gagging bills passed by Pitt and 14

15

16 17

For a discussion of Coleridge and science, see T. H. Levere, Poetry Realized in Nature: Samuel Taylor Coleridge and Early Nineteenth-Century Science (Cambridge: Cambridge University Press, 1981). Coleridge’s marginalia to Ørsted’s Ansicht are published in The Collected Works of Samuel Taylor Coleridge. Marginalia III, edited by H. J. Jackson and George Whalley (Princeton, NJ: Princeton University Press, 1992), pp. 995–1012; the annotated volume is in the British Library, C43 a17. H. Davy, Philosophical Magazine, 38 (1811), pp. 13–18. Coleridge, Marginalia (n. 2), p. 1001.

BETWEEN ENLIGHTENMENT AND ROMANTICISM

263

Grenville. Both Beddoes and Coleridge were sympathetic to the ideals of the French Revolution, which made each of them suspect in the eyes of government, and the subject of investigation by a government spy. In spite of being suspected of seditious purposes, Beddoes was a meliorist rather than a destroyer; he favoured improving the condition of life for the poor over fomenting revolution. Writing to his friend Davies Giddy, county sheriff of Cornwall and later, as Davies Gilbert, President of the Royal Society of London, Beddoes announced that he dreaded the Paris mob.18 Nonetheless, as an advocate of democracy, Beddoes seemed to the establishment as dangerous as social democrats would seem to McCarthyites almost two centuries later. Beddoes was in Bristol precisely because of his politics. Before settling in Clifton he had been reader in chemistry at Oxford, with the immediate prospect of a Regius chair in that subject. Government investigation of Beddoes had revealed what James Watt Jr., with admiring irony, called his “cloven Jacobin hoof,” and had led to correspondence with the Vice-Chancellor of the university, pointing out the unwisdom of nominating a crypto-Jacobin to a Regius chair. Beddoes thought that a democratic republic was the best form of government, as he wrote to Giddy in 1792: “I wish the republican spirit may now become the universal, as it is evidently the prevailing.”19 And again, in the fall of that year, “Vive l’egalité, vice G[od] S[ave] the K[ing].”20 The prospect of the Chair vanished, and although Beddoes suspected the reason, he was never formally told why. He did, however, find it wise to leave Oxford for Bristol.21 Politics apart, Beddoes was both a success and a pain while at Oxford. He was a great success as a lecturer—he boasted to Black that he had attracted “the largest class that has ever been seen at Oxford, at least within the memory of man, in any department of knowledge.”22 With the assistance of James Sadler, an instrument maker best known for his work on steam engines and later as the balloonist on Lord Macartney’s embassy to China, he was involved in carrying out a thorough overhaul of the chemical laboratory, and claimed that Sadler had constructed a gasometer that was a big improvement over Lavoisier’s.23 He was a pain to his politically conservative colleagues—a radical democrat in a tory institution would scarcely have been warmly welcomed; and he was a particular pain to Bodley’s Librarian.24 Beddoes rehearsed a litany of complaints, all of which may be familiar to today’s readers, even though we may be more fortunate in our librarians 18 19 20 21

22

23 24

Beddoes to Giddy 5 May 1791, Davies Giddy papers, Cornwall Record Office, Truro. Beddoes to Giddy 18 July 1792, Giddy papers 41/14. Beddoes to Giddy 19 November 1792, Giddy papers 41/38. For the background to these developments, see Dorothy A. Stansfield, Thomas Beddoes M.D. 1760–1808 (Dordrecht: Reidel, 1984); Roy Porter, Doctor of Society : Thomas Beddoes and the Sick Trade in Late-Enlightenment England (London and New York: Routledge, 1992); and T. H. Levere, Chemists and Chemistry in Nature and Society 1770–1878 (Aldershot, Hants., and Brookfield, Vt.: Ashgate [Variorum], 1994), essays V–VIII. Beddoes to Joseph Black 23 February 1788, Edinburgh University Library MS Gen 873/III/71,72. For background, see The History of the University of Oxford, general editor T.H. Aston, vol. 5, The eighteenth century, edited by L. S. Sutherland and L. G. Mitchell (Oxford: Clarendon Press, 1986). Beddoes to Joseph Black 15 April 1791, Edinburgh University Library MS Gen 873/III/200,201. [Thomas Beddoes], A Memorial Concerning the State of the Bodleian Library, and the Conduct of the Principal Librarian, Addressed to the Curators of that Library, by the Chemical Reader, 31 May [1787].

264

T. H. LEVERE

than Oxford seems to have been in the 1790s—inconveniently limited opening hours, inaccessible volumes, missing volumes, uneven acquisitions, incomplete runs of serials, inadequate staffing, and more besides. Beddoes was particularly concerned about the lack of care taken “to supply us with the authors of a country, who may justly contest the palm of Science and Literature, with those of any other nation.” He meant Germany. Beddoes acknowledged that there were few in Oxford who read German, but argued that if the “poets, divines, philosophers, and lawyers of Germany were within our reach, we should be tempted to study them.…We cannot surely be afraid, lest the labour of acquiring the language should be thrown away, unless we can suppose that the powers of Haller, Heyne, Meiners, and Michaelis, desert them, when they write in their mother-tongue.… But how can such writers, as…Doederlein,…Reimarus, Mendelssohn, or Lessing, be searched for new arguments…, while our high-priests of learning take no care to introduce their offerings into her temples?”25 Beddoes’s list is a striking one. Albrecht von Haller’s physiological and other writings made him essential reading for physicians, and some of his works, mostly published in Latin but a few in German, were also translated into Latin, French, English, and Italian. His presence on the list is no surprise; rather, his absence would have been remarkable. The Bodleian Library in 1787 had several of his works in Latin. Beddoes complained that the library had bought a copy of Haller’s Elementa Physiologiae with three volumes missing.26 “The other works of Haller are likewise incompleat. I find only…one [volume] of his Bibliotheca Anatomica:27 the Bibliotheca Med. Practicae is wanting, though a part of the work.”28 Christian Gottlob Heyne was a correspondent of Haller’s, one of the new breed of classical scholars at Göttingen who looked at the monuments of the past as well as texts, and whose work, mainly philological, explored Greek myths and the meanings of Greek art. Christoph Meiners, professor of philosophy at Göttingen, was also a literary historian and an historian of religion. His writing on the philosophy of aesthetics,29 hot off the press when Beddoes penned his complaint against Bodley’s Librarian, was a transitional work, recognizing the role of the senses, but also seeking “to attain the spiritual and active force of the ideal which shapes the material world in the first place.”30 Johann D. Michaelis had provided notes and some translation in the first German edition of the Oxford scholar Robert Lowth’s volume of lectures on sacred Hebrew poetry.31 Lowth’s book first appeared in English translation in 1787. Johann Christoph Doederlein 25 26

27

28 29 30

31

Ibid. pp.15–16. Albrecht von Haller, Elementa physiologiæ corporis humani. (Lausannæ, 1757–1766); the last three vols., 6–8, were subsequently acquired, but from the second edition (Lausanne, 1778). Haller, Bibliotheca anatomica, qua scripta ad anatomen et physiologiam facientia a rerum initiis recensentur, 2 vols. (Tiguri, 1774–1777) Beddoes, Memorial (1787), p. 10. Meiners, Grundriss der Theorie und Geschichte der schönen Wissenschaften (Lemgo, 1787). James Engell, The Creative Imagination: Enlightenment to Romanticism (Cambridge, MA: Harvard University Press, 1981), p. 110. Robert Lowth, De sacra Poesi Hebræorum prælectiones academicæ Oxonii habitæ.… Subjicitur Metricæ Harianæ brevis Confutatio: et Oratio Crewiana. Editio secunda,. accessionibus secundæ editionis Oxoniensis dilata. Notas et Epimetra adjecit J. D. Michaelis (Goettingæ, 1770).

BETWEEN ENLIGHTENMENT AND ROMANTICISM

265

was another scholar who translated and provided annotations and commentary to Hebrew scriptures.32 H. S. Reimarus was a deist of the Enlightenment, who argued that the rational human mind did not need revelation to arrive at the ideal religion. He did not publish his manuscript, Apologie oder Schutzschrift für die vernünftigen Verehrer Gottes (Apologia or Defence for the Rational Worshippers of God), but the playwright, critic, philosopher, and aesthetician Gotthold Ephraim Lessing did publish it in 1774–1778, after Reimarus’s death. And Moses Mendelssohn, Enlightenment rational Jewish philosopher, was the model for the central character in Lessing’s play Nathan der Weise of 1779. Enlightenment, rationality, idealism, theories of imagination, literature (including drama and ancient Hebrew poetry), literary criticism, the philosophy of aesthetics, the natural sciences, and medicine were all represented in the list that Beddoes set down to challenge the insularity and passivity of the Bodleian Library. He had constructed a virtual ideogram of the late Enlightenment as it merged into the early Romantic movement. Although it is fashionable now to follow the later Coleridge in seeing a fundamental incompatibility between the thought of the philosophes of the Enlightenment, and especially the French Enlightenment, and the idealism of the full-blown Romantic movement, there was a large measure of continuity in cultural life in the 1780s and 1790s, that represents a transition between extremes.33 Beddoes, in the range of his sympathies and his intellectual engagements, is himself a transitional figure, one whose enthusiasm for German literature, philosophy, and science was a powerful encouragement for Coleridge to visit Germany and immerse himself in the latest productions of German intellectual life. Beddoes helped Coleridge to engage with German Romanticism. Beddoes was remarkably well connected to the world of European science. His attendance at Joseph Black’s lectures at Edinburgh had given him the opportunity to meet other students from around the world. Black’s students came from Russia, France, Germany, Sweden, and numerous other countries.34 Beddoes translated a variety of foreign scientific and medical texts into English, including works by Spallanzani, Bergman, and Scheele, and from Latin and Spanish texts.35 It is impressive but not astonishing that Beddoes should have acquired the linguistic competence to read medical and chemical texts in a variety of European languages. Educated Englishmen would normally have been able to enjoy French and Italian operas, and knowledge of German was a reasonable response to the resurgence of German literature. Competent physicians still had Latin as their lingua franca. Swedish pre-eminence in mineralogy and metallurgy meant that it was advantageous for chemists to acquire familiarity with the 32

33 34

35

See, e.g. Salomons Prediger und hohes Lied. Neu übersezt mit kurzen erlaüternden Anmerkungen von D. Johann Christoph Do¨derlein (Jena, 1784). This is the main argument of Engell, The Creative Imagination (n. 30, above). See, e.g. T. H. Levere and G. L’E. Turner, eds., Discussing Chemistry and Steam: The Minutes of a Coffee House Philosophical Society 1780–1787 (Oxford: Oxford University Press, 2002), p. 9. A list of Beddoes’s publications, originally published in J. E. Stock, Memoirs of the Life of Thomas Beddoes M. D. (London and Bristol, 1811), is reprinted in Dorothy A. Stansfield, Thomas Beddoes M.D. 1760–1808 (Dordrecht: Reidel, 1984), pp. 282–284.

266

T. H. LEVERE

Swedish language.36 What was less to be expected was Beddoes’s engagement with German philosophy. The introduction of Kant into England was partly due to F. A. Nitsch, who lectured in London in 1794–1796, and who published A General and Introductory View of Kant’s Principles concerning Man, the World and the Deity, submitted to the Consideration of the Learned (London, 1796).37 But Beddoes had tackled Kant’s Kritik der reinen Vernunft no later than 1793. In his Observations on the Nature of Demonstrative Evidence, Beddoes discussed Kant’s system, observing that Kant “has raised to himself throughout Germany, a reputation superior to that of Wolff, and at least equal to that of Leibnitz”. He mentioned that Kant’s doctrines were making their way at the University of Göttingen. Beddoes at this date felt it necessary to take Kant seriously, but he had scarcely become a Kantian. He pointed out what he regarded as Kant’s errors, and argued that Hume had more accurately understood the nature of causation than had his German critic. He argued that even mathematics was based upon induction and observation, “whatever Mr. Kant may imagine.”38 But in the years that followed, Beddoes became increasingly sympathetic to Kant’s approach. In 1796 he wrote a positive review of Kant’s Zum ewigen Frieden39 for the Monthly Review, and in the Monthly Magazine for May 179640 expressed his admiration for the Kritik der Urtheilskraft (Berlin, 1790); he translated passages into English, urged the reader to tackle Kant for himself, and wrote a competent précis of Kant’s philosophy.41 1796 also saw Beddoes’s review of the third edition of Blumenbach’s anthropological essay, De generis humani varietati nativa (1795).42 Blumenbach was at the University of Göttingen, which, as we have seen, Beddoes had complimented on the excellence of its library, and had identified as one of the centres of learning where Kant’s philosophical system was making headway.43 Coleridge was to attend Blumenbach’s lectures. Beddoes prefixed his review of Blumenbach with an “Epistle” to Sir Joseph Banks, President of the Royal Society of London. Beddoes had already started the campaign for his Pneumatic 36

37 38

39

40 41

42 43

See, e.g. Brian Dolan, “Transferring Skill: Blowpipe Analysis in Sweden and England, 1750–1850,” in Brian P. Dolan, ed., Science Unbound: Geography Space and Discipline (Umeå: Umeå University Press, 1998), pp. 91–125; David Oldroyd, “A Note on the Status of A. F. Cronstedt’s Simple Earths and his Analytical Methods,” Isis 65 (1974), 506–512, and “Some Phlogistic Mineralogical Schemes, Illustrative of the Evolution of the Concept of ‘Earth’ in the 17th and 18th centuries”, Annals of Science 31 (1974), 269–305. For a general discussion, see René Wellek, Kant in England (1793–1838). Thomas Beddoes, Observations on the Nature of Demonstrative Evidence; with an Explanation of Certain Difficulties occurring in the Elements of Geometry; and Reflections on Language (London: J. Johnson, 1793), pp. 89–96. Immanuel Kant, Zum ewigen Frieden. Ein philosophischer Entwurf (Königsberg, 1795), reviewed in Monthly Review 20 (1796) (reviewer identified as Beddoes in Benjamin Christie Nangle, The Monthly Review, second series, 1790–1815 : indexes of contributors and articles_(Oxford : Clarendon Press, 1955) ). Monthly Magazine, or, British Register 1 (1796), 265–266. I was led to this review by Kathleen Coburn, ed., The Notebooks of Samuel Taylor Coleridge: Vol. 1 1794–1804, Bollingen Series L (New York: Pantheon Books, 1957), 249n, which erroneously places the review in the Monthly Review. Monthly Review 21 (2) (1796), p. 515. For the role of Kant’s Kritik der Urteilskraft in the development of Blumenbach’s physiology and beyond, see J. Steigerwald, “Lebenskraft in reflection: German perspectives of the late 18th and early 19th centuries,” Ph.D. dissertation, University of London, 1998.

BETWEEN ENLIGHTENMENT AND ROMANTICISM

267

Institution; and it was clear that Banks’s endorsem*nt, if only it were forthcoming, would encourage the medical establishment, which included numerous Fellows of the Royal Society, to support Beddoes’s campaign. This was labour wasted. Banks, in spite of vigorous lobbying by Georgiana Cavendish, duch*ess of Devonshire, was unwilling to give Beddoes any encouragement. Banks argued that his objection was based upon the probable failure of the research and therapeutic programs of Beddoes’s institution, but Beddoes’s friends were convinced that the real reason was political, and that Banks had caught a glimpse of Beddoes’s Jacobinism.44 Politics, in the form of responses to the French Revolution and the Terror that followed, were a constant subtext and often the main text in the 1790s. Beddoes while at Oxford had waxed eloquent in his correspondence about the advantages of a Republic, “the only form of government consistent with honesty and common sense.”45 It was no surprise that he had had to leave the Tory bastion of Oxford. In 1794, soon after Beddoes’s departure for Clifton, Louis had been executed, England was at war with France, and the whole country was agitated by fears of invasion, food shortages, rumours of sedition, and, for a minority, uneasy attempts to maintain civil liberties, and the possibility of democracy. Bristol, a major port, looked across the sea to Ireland, and feared that the French might find support in that country against England, and might use it as a springboard for invasion. On 2 January 1794, Henry Dundas, in charge of the war and also of security at home, learned that a French dancing master and a miniature painter in Bristol were both disaffected, clear evidence of the decadence and danger represented by French culture. Dundas contemplated further restrictions on the limits within which aliens might reside.46 Only a minority was opposed to continuing the war with France. Beddoes was one of that minority, and published a pamphlet addressed to the inhabitants of Bristol, asking “Where would be the harm of a speedy peace?”47 The paranoia that saw danger in a French dancing master was reflected in increased restrictions on free speech. Bills were proposed in parliament to suspend habeas corpus and restrict the right of assembly. Beddoes, who had already been suspected of sedition, was prominent in opposing what he was the first to call “gagging bills,”48 and published a pamphlet, “A Word in Defence of the Bill or Rights.”49 A public meeting was held in Bristol’s Guildhall, to oppose the gagging

44

45

46 47 48

49

T. H. Levere, “Dr. Thomas Beddoes and the establishment of his Pneumatic Institution: A tale of three presidents,” Notes and Records of the Royal Society of London 32 (1977), pp. 41–49 at 44. Beddoes to Davies Giddy 8 November [1792], Cornwall Record Office, Truro, Davies Giddy papers DG 41/5. Dundas to the Mayor of Bristol, 2 January 1794, Public Record Office, Home Office papers 43/4. Thomas Beddoes, “Where would be the harm of a speedy peace?” (Bristol: N. Biggs, 1795). Lewis Patton, ed., The Watchman in The Collected Works of Samuel Taylor Coleridge (London and Princeton: Routledge and Kegan Paul, and Princeton University Press, 1970), p. 344n. Beddoes, “A Word…” (Bristol, 17 November 1795), reprinted in Lewis Patton and Peter Mann, eds., Lectures 1795 On Politics and Religion in The Collected Works of Samuel Taylor Coleridge (London and Princeton: Routledge and Kegan Paul, and Princeton University Press, 1971), pp. 371–384. Coleridge reprinted an “Extract from Dr. Beddoes’s POSTSCRIPT to his Defence of the BILL of RIGHTS against GAGGING BILLS” in The Watchman no. 10, 13 May 1796; it may be found in Lewis Patton, ed., The Watchman in The Collected Works of Samuel Taylor Coleridge (London and Princeton: Routledge and Kegan Paul, and Princeton University Press, 1970), pp. 344–346.

268

T. H. LEVERE

bills. Beddoes argued that it was only because of unexpected opposition that Pitt had left out the clause originally restricting “in private houses conversations on political subjects to a certain number of persons.” But freedom of speech was still under siege, and “since the man who wages war without success, cannot expect to make peace without disgrace, he will cry out for a prolongation of hostilities.”50 One of William Wordsworth’s correspondents was at the meeting, and reported to him: “Dr. Beddoes is no Orator, but spoke to the purpose.”51 Coleridge was at the meeting, and also spoke. If Coleridge and Beddoes had not met before, they met then, and Coleridge swiftly became one of Beddoes’s warmest admirers. Politics may well have brought them together in 1795, but their intellectual engagement was much broader, and of lasting consequence. Beddoes read widely and reviewed widely. His library was impressive, as was his familiarity with foreign and especially German literature. When a Dr. Frank of Vienna visited Beddoes some years later (in 1803), he was kept waiting for a while. Then Beddoes entered the room bearing an armload of books by different Dr. Franks, asking, “Which Dr. Frank are you?” There were three distinguished Dr. Franks from Vienna alone, the cousins Joseph and Ludwig, and Joseph’s father Johann. The Frank visiting Beddoes turned out to be Joseph, who subsequently remarked that Beddoes “reads German as well as he does English and is intimately acquainted with all our best authors.”52 Joseph and Ludwig were to write on the doctrine of the Edinburgh physician John Brown, whose work Beddoes had edited.53 Coleridge appears soon to have been given access to Beddoes’s library.54 In 1801, he wrote to Davy: “I received a letter this evening from Dr. Beddoes who immediately wants the Books in the inclosed Parcel.”55 It is not clear from the context whether Coleridge was overdue in returning borrowed books, a fault of which he was occasionally guilty,56 or whether he was sending Beddoes books that Beddoes had ordered through him from London. He was reading a variety of mainly English scientific writers in 1795 and 1796, and one year later proposed his cheerfully preposterous but subsequently more than half achieved plan of studies: “I should not think of devoting less than 20 years to an Epic Poem. Ten to collect materials and warm my mind with universal science. I would be a tolerable 50 51

52

53

54

55

56

Ibid. pp. 2–5. Azariah Pinney to William Wordsworth 26 November 1795, quoted in Lectures 1795 On Politics and Religion_(1971) p. xlv. John Edmonds Stock, Memoirs of the life of Thomas Beddoes, M.D., with an analytical account of his writings (London: John Murray, 1811.), p. 301. John Brown, The Elements of Medicine … Translated from the Latin, with comments and illustrations, by the author. A new edition, revised and corrected. With a biographical preface by Thomas Beddoes, M.D., and a head of the author, etc., 2 vols. (J. Johnson: London, 1795). Coleridge was much interested in Brunonian medicine, as shown by notebook entries while he was in Germany. See, e.g. Kathleen Coburn, ed., The Notebooks of Samuel Taylor Coleridge: Vol. 1 1794–1804, Bollingen Series L (New York: Pantheon Books, 1957), entries 388–389 and nn. George Whalley, “The Bristol Library Transactions of Southey and Coleridge, 1793–8,” The Library, Transactions of the Bibliographical Society, 5th Series, 4 [1949], 114–131. Coleridge to Humphry Davy, 4 May 1801, in Earl Leslie Griggs, ed., Collected Letters of Samuel Taylor Coleridge. Volume II 1801–1806 (Oxford: Clarendon Press, 1956), p. 726. See, e.g. Coleridge to John Hookham Frere, 16 July 1816, in Collected Letters of Samuel Taylor Coleridge. Volume IV 1815–1819 (Oxford: Clarendon Press, 1959), p. 656.

BETWEEN ENLIGHTENMENT AND ROMANTICISM

269

Mathematician,”—he never was,—“I would thoroughly know Mechanics, Hydrostatics, Optics, and Astronomy, Botany, Metallurgy, Fossilism, Chemistry, Geology, Anatomy, Medicine—then the mind of man—then the minds of men—in all Travels, Voyages and Histories.”57 Coleridge could not afford to buy scientific journals, but he used Beddoes’s library and the Bristol Library Society.58 Beddoes encouraged him to tackle German as well as English sources, and Coleridge began to study German by the spring of 1796. He proposed getting a London bookseller to publish a translation of Schiller’s complete works, at the trifling cost of sending Coleridge and his wife to Jena. “If I could realize this scheme, I should there study Chemistry & Anatomy, [and] bring over with me all the works of Semler & Michaelis, the German Theologians, & of Kant, the great german Metaphysician.” It was either that, or becoming a dissenting minister.59 Coleridge, of course, did go to Germany, although without his wife. He ended up in Göttingen, and not in Jena, where Schiller resided. He did return to England with boxes of German books, a fluent reading knowledge of German, and an excruciating accent in that language. Although he progressed in later years from Kant to Schelling, and then rejected Schelling’s philosophy because he considered that it made nature absolute, he never ceased to be shaped by the programme of reading that Beddoes had first catalysed. He seems not to have studied chemistry in Göttingen, although he studied a good deal of natural science, including comparative anatomy and geology. When he returned from Germany, he headed for Bristol, and an habitué of Beddoes’s now functioning Pneumatic Institution. Young Humphry Davy, brilliant, charismatic, a poet as well as a chemical philosopher in the making, was working as the chemist in that institution, and he and Coleridge soon became friends and inspirations to one another. They both tried breathing nitrous oxide, Coleridge attended some of Davy’s work with patients at the Institution, and they spurred one another on to glory. In 1801 Davy left Clifton for the Royal Institution in London. In 1804 Coleridge left England for Malta, in hopes of recovering his health, much battered by opium addiction. He was remarkably active for a sick man. In 1805 he moved on to Naples and Rome, and in 1806 returned to England and to his friends and family. Two years later, Beddoes died at the early age of 48, a disappointed and lonely man.60 Coleridge was devastated. “I have received a very severe and very abrupt Blow in the Death of Dr. Beddoes—he was good and beneficent to all men, but to me he was tender and affectionate. Few Events have taken so much Hope from my Life.”61 Beddoes had been his friend, his mentor, and his physician, and Coleridge missed him sorely. Beddoes’s library was sold at auction by Leigh and Sotheby. The sale occupied ten days, beginning on 10 November 1809, and his library was described in the catalogue

57

58 59 60 61

Coleridge to Joseph Cottle, [early April 1797], in Collected Letters of Samuel Taylor Coleridge. Volume I 1785–1800 (Oxford: Clarendon Press, 1956), pp. 318–319. Whalley [1949]. Coleridge to Joseph Cottle, Collected Letters vol. I, pp. 209–210. Stansfield (1984), p. 1. Coleridge to Daniel Stuart, [3 January 1809], in Collected Letters of Samuel Taylor Coleridge. Volume III 1807–1814 (Oxford: Clarendon Press, 1959), p. 160.

270

T. H. LEVERE

as “containing a very capital collection of modern publications in all the departments of surgery and medicine; voyages and travels, antiquities, natural history, and belles lettres. Likewise all the best German writers on the above subjects.”62 The sale of the German books alone took the first five days of the sale. The catalogue listed 1,071 lots of German titles, which constituted just over half the entire library. Medical texts were of course the largest single group, including seven volumes of the writings of three Drs. Frank.63 John Brown’s system, with its notion of excitability, was one with which Beddoes had experimented over the years, with some ambivalence, preserving it in his practice of medicine while subjecting it to theoretical criticism. In 1795 he had provided an introduction and notes to a new edition of Brown’s Elements of Medicine, as a service to Brown’s widow.64 Brunonian medicine became widely known in Germany by the time that Coleridge turned up in Göttingen, where he read a long review of works on that system spread over twelve numbers of the Allgemeine Lietratur-Zeitung (Jena and Leipzig, 11–20 February, 1799). The subject obviously aroused lively controversy in Göttingen, for in 1802, when Coleridge was back in England, the cavalry was called out to put down rioting between Brunonians and their critics.65 Beddoes possessed, besides numerous works in other languages, several of the German texts on Brown’s system.66 He had numerous works relating to electricity, medical electricity, and galvanism, and his German volumes67 included those of the Naturphilosoph Johann Wilhelm Ritter,68 whom Davy deplored.

62

63

64

65 66

67

68

A Catalogue of the Very Valuable and Extensive Library of Thomas Beddoes, M.D. of Clifton, near Bristol, lately deceased.… which will be sold by auction, by Leigh and S. Sot heby, booksel l er s, at their House, No. 145, Strand, opposite Catherine Street, On Fr iday, November 10, 1809, and Nine following Days, (Sundays excepted) at 12 o’Clock_[np, nd], 71 pp. The auctioneer’s copy is listed in the British Library catalogue as Sotheby 60 (2), but has gone astray during or since the move to the new building. I obtained a microfilm of this copy some years ago. Johann Peter Frank, System einer Medizinischen Polizey, 4 vols. (Manheim, 1784–1788); J. S. Frank, Versuch einer Arzneymittellehre nach Grundsaetzen der Erregungs theorie (Wien, 1803); J. S. Frank, Erläuterungen der Erregungs theorie (Heilbron, 1803); Joseph Frank, Grundriss der Pathologie, nach den Erregungs theorie (Wien, 1803). John Brown, The Elements of Medicine, translated by the author, new edition with biographical preface and introduction by T. Beddoes, 2 vols. (London, 1795). The context is discussed in T. H. Levere, “Dr. Thomas Beddoes: The interaction of pneumatic and preventive medicine with chemistry,” Interdisciplinary Science Reviews, 7 (1982), 137–147. Levere, Poetry Realized in Nature (1981), pp. 202–203 and notes. Cappel, Beurtheilung des Brownischen Systems der Medecin (Göttingen, 1800); Cappel, Beurtheilung des Brownischen Systems (Göttingen, 1797); M. Detten, Vorschlag zur Brownisirung des Organismus (Munster, 1800); Embirn, Pharmacologia Browniana oder Handbuch der Heilmittel (Stuttgart, 1793); von Hogen, Vorzuge der Brownschen Praxis (Ludwigsburg, 1803); A. F. Marcus, Prüfung des Brownschen Systems (Weimar, 1797); A. F. Marcus, Prufung des Brownischen Systems, 2 vols. (Weimar, 1798): C. H. Pfaff, John Brown’s System der Heilkunde (Copenhagen, 1796); A. Röschlaub, Von dem Einflusse der Brownschen Theorie (Wurzburg, 1798). F. L. Augustin, Vom Galvanismus und dessen Medicinischer Anwendung (Berlin, 1801); E. A. Eschke, Galvanische Versuche_(Berlin, 1803); Hellwag & Jacobi, Erfahrungen über die Heilkraefte des Galvanismus (Hamburg, 1802); C. H. Pfaff, Über Thierische Elektricitaet (Leipzig, 1795); F. Pilger, Versuch über den Galvanismus (Giessen & Darmstadt, 1801); C. A. Struve, System der Medicinischen Electriciataets Lehre, 2 vols. in 1, with (Breslau, 1802). J. W. Ritter, Galvanismus (Weimar, 1798); Beytraege zur Kenniss des Galvanismus, 2 vols. in 5 parts (Jena, 1800); System der Körper (Leipzig, 1805).

BETWEEN ENLIGHTENMENT AND ROMANTICISM

271

Davy had attributed Ritter’s errors to the influence of Kant; Schelling was in fact more directly the culprit, but Davy read neither Kant nor Schelling. Beddoes, however, as we have seen, both read and reviewed Kant, and owned nine of his books, including the dynamic treatise on the metaphysics of natural science, in the edition that Coleridge began to annotate in ca. 1806.69 In England in the 1790s and 1800s, such a collection of the works of Kant is remarkable; even more remarkable is Beddoes’s possession of Friedrich von Schelling’s Erster Entwurf eines Systems der Naturphilosophie (Jena and Leipzig, 1799), and Johann Gottlieb Fichte’s Das System der Sittenlehre nach den Principien der Wissenschaftslehre (Leipzig 1798)—both works also annotated by Coleridge. By now, it comes as no surprise that the scientific works, by Blumenbach and others, read by Coleridge in Germany and on his return, were, for the most part, also in Beddoes’s library. There is so much in Beddoes’s library that much of Coleridge’s German reading was bound to involve the same titles, whether or not he read them in Beddoes’s collection; and he had also made good use of the library in Göttingen.70 But Coleridge did not go to Germany with the intention of making a study of the natural sciences. He went to learn philosophy, and with the intention, minimally realized, of translating Schiller into English.71 A key work in this project of translation was Die Räuber (The Robbers) (Leipzig, 1782), which Beddoes possessed, along with seven other volumes of Schiller’s works. Along with Schiller, in the sale catalogue of Beddoes’s library that came hard upon his death, we find 2 volumes of Novalis, 5 volumes of Mendelssohn, 6 volumes of Goethe, 20 volumes of Herder, 28 volumes of Jean Paul, and 35 volumes of Lessing, to name a few. Beddoes had a real relish for the best and most recent of German writing, in literature, philosophy, science, and medicine. He reviewed as well as read, and his conversation, to those who were not distressed by democratic politics, was as stimulating as it was wide-ranging. During the sale of Beddoes’s library, his books were sold in 2,131 lots. I have merely begun to use the sale catalogue. The next step will be to identify the purchasers of Beddoes’s books. The auctioneer or his assistant wrote down the buyer for each lot—surnames only, occasionally with a title (mostly Dr.). Perhaps two-thirds of the names of buyers of the 1,071 lots of German books were themselves German, and one-third English; the émigré community, although active, was clearly not alone in ensuring that there was an appetite for German

69

70

71

Kant, Kritik der reinenVernunft (Riga 1788); Anthropologie in pragmatischer Hinsicht abgefasst (Königsberg, 1798); Der Streit der Facultäten, in drey Abschnitten (Königsberg, 1798); Prolegomena zu einer jeden künftigen Metaphysik die als Wissenschaft wird_auftreten können (Riga, 1783); Metaphysische Anfangsgründe der Naturwissenschaft, Zweyte Auflage (Riga, 1787); Versuch den Begriff der negativen Grössen in die Weltweisheit einzuführen (Königsberg 1763); Beobachtungenüber das Gefühl des Schönen und Erhabenen (Riga 1771); Grundlegen zur Metaphysik der Sitten (Riga, 1786); Lebendigen Kräfte (Königsberg, 1746). For Coleridge’s marginalia, see Coleridge, Marginalia III (n. 15 above), conveniently in the same volume as his marginalia to Ørsted. A. D. Snyder, “Books Borrowed by Coleridge from the Library of the University of Göttingen, 1799,” Modern Philology 25 (1928), 377–380. For Coleridge and Schiller, see The Collected Works of Samuel Taylor Coleridge: 16. Poetical Works, edited by J. C. C. Mays, 6 vols. (I–III in 6 parts) (Princeton, NJ: Princeton University Press, 2001).

272

T. H. LEVERE

books in London in 1808. Beddoes could not always have found it easy to acquire German books—faced with the likely difficulties during the Revolutionary and Napoleonic Wars, his achievement is remarkable, and the sale of his library represented a real opportunity for those thirsty for German culture. We know from Coleridge that there were few avenues in London that afforded him access to the latest German books. Beddoes’s library, right up to his death in 1808, offered an extraordinarily rich resource. We do not know which of the German books that Coleridge read came from that library, and in most cases we are unlikely ever to know; but at least in a few cases, I hope to find out. University of Toronto

ØRSTED’S PRESENTATION OF OTHERS’—AND HIS OWN—WORK KENNETH L. CANEVA

Abstract Ørsted is best known for his discovery of electromagnetism in 1820, clearly his most important and arguably h is only significant contribution to ‘science’ understood as a body of knowledge pertaining to a canonical set of fundamental phenomena. Yet if by ‘science’ one understands the flesh-and-blood, words-on-paper world of ongoing human activity that produces such knowledge, then Ørsted’s contemporary significance looms much larger. It turns out that a striking amount of the work of other scientists became known principally through his agency, sometimes becoming so transformed in the process that what the larger world of science took, and continues to take, as others’ work in fact bears the strong stamp of Ørsted’s transmission. His fluency in German and French, his extensive European travels, and his wide circle of acquaintances all facilitated his personal influence. This paper examines in detail Ørsted’s role in presenting—and significantly transforming—Winterl’s dualistic chemistry, in bringing before a Parisian audience Ritter’s experiments on the chemical action of light and his construction of a storage column for galvanic electricity, and in recasting Seebeck’s discovery of the “magnetic polarization of metals by temperature difference” in terms of his conception of thermoelectricity. It further examines some of the changes Ørsted made in the presentation of his own work to different audiences—changes that throw in particular relief Ørsted’s complex relationship to German Naturphilosophie—as it identifies some of the peculiar aspects of his presentation of his electrodynamical work of 1820. The result is an enhanced appreciation of Ørsted’s important place in the science of his day, a deeper understanding of Ørsted’s own scientific work, and an illustration of the complex interrelationship between individuals and the scientific community in the creation of generally accepted scientific knowledge.

1. INTRODUCTION Hans Christian Ørsted (1777–1851) is best known for his discovery of electromagnetism in 1820, clearly his most important and arguably his only significant contribution to ‘science’ understood as a body of knowledge pertaining 273 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 273–338. © 2007 Springer.

274

K. L. CANEVA

to a canonical set of fundamental phenomena.1 Yet if by ‘science’ one understands the flesh-and-blood, words-on-paper world of ongoing human activity that produces such knowledge, then Ørsted’s contemporary significance looms much larger. It turns out that a striking amount of the work of other scientists became known principally through his agency, sometimes becoming so transformed in the process that what the larger world of science took,– and continues to take,– as others’ work in fact bears the strong stamp of Ørsted’s transmission. The most important example of this is surely his thermoelectric interpretation of what Seebeck had seen as the magnetic polarization of metals produced by a temperature difference. Just as it was Ørsted’s interpretation that determined (in particular) French scientists’ perception of Seebeck’s work, it is Ørsted’s conceptualization and terminology that have survived. Of more temporally and geographically localized significance was Ørsted’s campaign on behalf of Winterl’s chemical theory. Then as now what many people have understood as Winterl’s chemistry was what Ørsted chose to emphasize, often—and significantly—in language quite foreign to the tenor of Winterl’s appreciably less accessible Latin. And it was through Ørsted’s intercession that important aspects of Ritter’s researches became more generally known. Indeed, since none of Ritter’s own papers were translated into French, it was primarily through Ørsted’s reports in Parisian journals and his on-the-spot reportings both to individual scientists and before the most important scientific societies in Paris that French-speaking scientists became acquainted with Ritter’s work. It was precisely through that reporting, moreover, that Ørsted first established a Parisian reputation, not as a result of his own then-modest scientific attainments. More generally, Ørsted’s contemporary visibility owed a great deal to a succession of instances in which he played a dominant role in bringing to attention the work of other scientists. Ørsted was in fact uniquely well situated to perform that important role. As a well-traveled Dane fluent from an early age in German, well-grounded in the academic Latin of his age, and eager to perfect his French during repeated long stays in the French capital, he moved with ease among the principal Continental scientists of his day, an amazing number of whom he met personally. Having essentially unimpeded access to the relevant literature in all the major languages— he also knew English—he was at the same time acutely aware of profound differences in style between (most importantly) German and French scientists. It is thus not surprising that Ørsted regularly adapted his presentation not only of others’ but also of his own work depending on which audience he was writing for. In addition, changes in Ørsted’s own views over the 50-year span considered here prompted him to change the way he presented himself to the world. In particular, the way in which he announced his discovery of electromagnetism and the language he consistently used to represent it offer important insights into his conception of science.

1

I use double quotes (“…”) to mark actual quotations, single quotes (‘…’) as ‘scare quotes’ to call attention to words used in some special way.

ØRSTED’S PRESENTATION OF OTHERS’

275

I. ØRSTED’S PRESENTATION OF OTHERS’ WORK 2. ØRSTED’S PRESENTATION OF WINTERL’S CHEMISTRY Sometime after 25 July 1800—the date of its preface—there appeared in Buda a Latin treatise by Jakob Johann Winterl (1732–1809), the obscure GermanHungarian professor of chemistry and botany at the University of Pest, entitled Prolusions toward a Chemistry of the Nineteenth Century (hereafter Prolusiones), of which its author sent more than 70 copies to academies, journal editors, and individual physicists and chemists.2 One of its earliest readers, and surely its most enthusiastic, was the young Ørsted, who in August 1801 had embarked on an ambitious journey, lasting until January 1804, to visit many of the leading scientists of Germany, France, Belgium, and Holland.3 Soon after his arrival in Berlin he wrote to Ludvig Manthey (1769–1842), a pharmacist and chemistry teacher at the Academy of Surgery, of the enthusiasm he had been nurturing for some weeks: I’m reading no book with greater diligence or pleasure in my evening hours than Winterl’s Prolusiones ad chemiam seculi decimi noni. I find with every new perusal more harmony and genius in it. I long only to repeat some of his principal experiments; but the terrible prejudice—I dare to call it such, since I am as yet acquainted with no one who has censured him in such a way that one could hear that he had read him—that prevails here and everywhere against it makes me somewhat wary of even talking about it. But I hope the occasion shall arise. I convinced Ritter to read it and had the pleasure of hearing his judgement agree with mine. Be all that as it may,…I more and more find signs of truth everywhere in this somewhat strange book.4

For the next two months or so Ørsted sought hard to stimulate interest in and support for Winterl’s ideas among Berlin scientists, encountering mostly misunderstanding and criticism from those who often had not even read the work, its being in Latin apparently an appreciable obstacle to many. He reported discussing it with polymath physical scientists Ernst Friedrich Wrede (1766–1826) and Paul Louis Simon (1767–1815), chemists Valentin Rose (1762–1807) and Jeremias Benjamin Richter (1762–1807), and physicist Paul Erman (1764–1851), while Rose and Erman urged him to prepare an exposition of Winterl’s ideas for presentation before the Berlin Philomatische Gesellschaft, which he appears to

2

3

4

Winterl 1800, xi; Ørsted 1803a, 136, n. 1. In a letter to Ørsted of 12 April 1803, Winterl recorded his disappointment over the lack of response from the Gelehrte for whom he’d given his Viennese bookseller 118 copies for distribution (Ørsted 1920b, 2, 603). The Paris Academy received its copy on 1 Nivôse, an XI = 22 December 1802 (PV, 2 [1912], 605]). See the travel diary and letters published (often abridged) in Ørsted 1870, 1, pp. 15–180. This extremely informative source deserves to be edited in full and translated into English. Ørsted, letter of 4 December 1801 to Manthey, in Ørsted 1870, 1, p. 30. Ørsted had gone to see Ritter in Weimar on 18 September 1801, beginning what was to be an extended and intensive relationship (travel diary, entry of 31 August 1801, in ibid., p. 26). Seven months later, Manthey wrote that he had read the book but did not share Ørsted’s enthusiasm for it (letter to Ørsted of 17 July 1802, in ibid., p. 75). Unless otherwise noted, all translations are my own and all italics are as in the original. I thank David Wharton for help with some of the Latin passages. I have (usually) silently omitted internal cross references (e.g. to other numbered paragraphs) from quoted passages, and have not reproduced the regular italicization of people’s names.

276

K. L. CANEVA

have done in February 1802.5 At some point during his five-and-a-half-month sojourn in Berlin—from the end of November 1801 till the middle of May 1802— he performed experiments in the laboratory of Sigismund Friedrich Hermbstädt (1762–1833) which agreed with Winterl’s theory of acids.6 A year later, in Paris, he spoke with Louis-Bernard Guyton de Morveau’s (1737–1816) assistant, CharlesBernard Desormes (1777–1862), about Guyton’s examination of Winterl’s chemistry, reporting that three “very bad” experiments had been performed that proved nothing. Ørsted was pessimistic: “I predict that Winterl will be condemned.”7 For the rest, Ørsted’s chief efforts on Winterl’s behalf were directed at German-speaking chemists, where personal contacts played an important role. Ørsted’s energetic campaign on Winterl’s behalf culminated in the publication in 1803 of his Materials for a Chemistry of the Nineteenth Century (hereafter Materialien), a book he had finished during a period of intense collaboration with Ritter in Jena between 13 August and 4 September 1802.8 As he noted in his thirdperson autobiography with regard to the period of great scientific ferment during the early years of the century following the advent of “the newer [i.e. dynamical] philosophy” and Volta’s discovery of his pile, “Winterl had published his Prolusiones in chemiam seculi decimi noni, a book full of grand ideas, but the reading of which one had found so little attractive that the young Danish traveler can be said in a certain fashion to be the one who introduced the Hungarian chemist among the Germans.”9 Indeed, to this day it appears that most people who have any notions at all of Winterl’s chemistry have derived them more from Ørsted’s exposition than from the little-read Latin original. Since my task is precisely to assess Ørsted’s presentation of Winterl’s work, the first order of business must be to take stock of just what it was that Winterl said, whereby I will be guided by the fact that ultimately it is Ørsted’s work and not Winterl’s that centrally concerns us here. Without trying to give a full summary of the book’s contents, I nevertheless wish to convey the tone and language of the work, especially since there appears to be no modern exposition of Winterl’s work based on a direct reading of the Prolusiones.10 The principal subject of Winterl’s Prolusiones is the nature of acidity and, by extension, of basicity: the “First Prolusion on the Cause of Acidity” occupies the first 168 pages, the “Second Prolusion on the Substrate of Azotic Air,” in which

5

6 7

8 9 10

Ørsted, travel diary, entries for 31 December 1801 and 10 January 1802, and letters of 7 and 8 February to Anders Sandøe Ørsted (his brother) and Manthey, in Ørsted 1870, 1, pp. 31, 36, 46–47, 51–52. In the third of these he reported that “except for Ritter, to whose attention I called it, I am acquainted with no chemist or physicist, either in Berlin or elsewhere, who even knew anything decent about its contents.” Wrede’s given names are also reported as Erhard Georg Friedrich. Ørsted 1804b, 322 = 1920a, 1, p. 211. Ørsted, travel diary, entry for 26 March 1803, in Ørsted 1870, 1, p. 128; cf. Jacobsen 2000, pp. 105–106, 2001, p. 202. Ørsted, travel diary, in Ørsted 1870, 1, p. 76. Ørsted 1828, p. 521. Despite the fact that he never cites directly from the Prolusiones, Snelders (1970, pp. 231–233) provides a good brief summary of Winterl’s basic ideas. Szökelfalvi-Nagy (1971, pp. 41–44) also provides a useful sketch of Winterl’s ideas, though one which is based solely on works published after the Prolusiones, and which fails to distinguish Ørsted’s interpretations from Winterl’s original ideas.

ØRSTED’S PRESENTATION OF OTHERS’

277

he introduced the notorious and ill-fated andronia, the last 83.11 Andronia was a ponderable, isolatable acidic substance, a new earth (terra) implicated in a welter of diverse phenomena.12 Alas for Winterl (and Ørsted), most contemporaries devoted their critical attention to the unsupportable experimental claims concerning andronia, largely ignoring the more important theoretical issues concerning acid-base reactions and Winterl’s idiosyncratic chemical ontology. Suggesting (in the preface) that chemists’ rejection of the doctrine of phlogiston had induced them too hastily to accept all of Lavoisier’s new chemistry, Winterl reported that his own experiments—in particular, those dealing with the cause of acidity—had prevented him from supporting many of Lavoisier’s opinions.13 [T]he more I worked in this arena the weaker the arguments became that establish vital air as the source of acidity; on the contrary, those arguments that affirm a specific principle of acidity, deriving from glowing heat [Caloricum lucidum] and a certain electrical state of bodies, became stronger; and at length all doubt disappeared when I recognized that vital air itself is only an acid sui generis… ; when I saw the diminished acidity in neutralizations—with the opposite principle coming to them from basic bodies— result in heat [Caloricum]; [when I saw] the constituent parts of heat moreover again be weakened [and] split apart by [the action of] light and be expended in oxidizing and basifying suitable substrates; when at length I felt I had come to where I might distinctly see all attractions, secretions, decompositions, and recombinations, by which the greatest as well as the least changes are performed in nature, to be effected solely by the components of heat and by light.14

That was about as much as Winterl explicitly criticized Lavoisier’s chemistry. Much later he expanded upon his views on heat and electricity, to which we shall return anon. After recounting some of the personal prehistory behind his work, including his overture to the Göttingische Gesellschaft der Wissenschaften, and urging an exchange of publications with foreign scientists, he closed the preface with a shot across the bow of the dynamists then in apparent ascendency: “I anticipate one kind of future readers whom this work will displease because I will posit matter (the substrate) to be destitute of forces, against the unaddressed and (as they will think) unexamined system they call dynamical, and I will establish a material cause of attraction.”15 Indeed, the concept of force played only an occasional role in Winterl’s schema. Notwithstanding his intention to give attraction (and repulsion) a material underpinning, his most common—though still relatively infrequent—application of the term force was to the forces of attraction and repulsion—he spoke of vires attrahentes reciprocæ and vires renitentes attractioni, “reciprocal attracting forces” and “forces resisting attraction”—and to the magnetic force.16 He also occasionally spoke in a general way now 11

12

13 14

15

16

In 1799 Winterl had sent a copy of the first part, entitled (in the conventional style) Experimenta et observationes de causa aciditatis, to the Königliche Gesellschaft der Wissenschaften in Göttingen, which published an extensive summary of its experimental portion in the Göttingische Anzeigen von gelehrten Sachen (Winterl 1800, vi; Anonymous 1800). Winterl (1800, p. 170) explained in a footnote that “I chose the name andronia since one might discover another new earth (thelyca), very similar to andronia in that it strives to unite itself with many bodies, but different with regard to its basic quality (as though with regard to sex).” He expected his readers to recognize the Greek roots andro—(male) and thely—(female). Winterl 1800, ii–iii. Winterl 1800, iii–iv. When quoting originally inflected words in isolation from their grammatical context, I have reduced them to standard nominative form. Winterl 1800, xi. Although I haven’t found any explicit reference to atoms, his occasional reference to the (obviously very small) “parts” of the substrate and of bodies suggests a leaning in that direction (130, 140). Winterl 1800, pp. 3–5, 155. I have not tried to record every usage.

278

K. L. CANEVA of the “force” of an acid, now of the “force of a base” in drawing the principle of acidity away from its substrate, or more particularly of an acid’s “force of blunting bases” and a base’s “force of blunting acids.”17

Winterl began the body of his work by criticizing the various ways in which people had sought to characterize acids and bases, for example by their taste, by the change in color they produce with vegetable pigments, or by the neutralization of bases by acids and acids by bases (“Basium per Acida, Acidorum per Bases solubilitas”).18 His solution, which he introduced gradually with many particulars, was to assign “principles” of acidity and basicity to different substances as the underlying “animating” causes of acid-base properties controlling their reactions.19 Winterl was thus continuing the tradition that sought to explain the common properties of a class of substances under the assumption that they must all contain the same quality-bearing yet substantial principle. As substances sui generis—not forces—they retained their identity and—potentially—their power to act even when their effects were blunted as a result of the interaction of the material substances that were, so to speak, their carriers.20 In an early application of these ideas he argued that spirit of wine itself has the power of blunting acids [virtus obtundendi Acida], but for a very different reason than true bases: the latter blunt the principle of acidity [Principium aciditatis] by means of a certain contrary basic principle, in such a way that it does not manifest itself by any sensible quality either in the salt generated or outside thereof; but alcohol separates the principle of acidity from acids in such a way that this principle adheres unchanged partly to the alcohol itself, partly to contiguous bodies by an induced or increased degree of acidity, as will be more fully shown in §15. Accordingly, it is seen that blunting [obtusatio] does not pertain solely to the residuum; rather, more strictly, the reciprocal blunting of the very principles animating acids and bases [ipsa Principia Acida atque Bases animantia] pertains to their more intimate nature.”21

17

18

19

20

21

Winterl 1800, pp. 6, 32–33 (vis Baseos), 77 (vis obtundendi Bases and vis obtundendi Acida). Winterl’s regular and frequent terms for the reciprocal action of acids and bases were obtundere (as a verb) and obtusatio (as a noun), the basic meaning being “to blunt or dull,” not “to annihilate the effect of.” I don’t find him to have used the corresponding adjective, obtusus (“blunted, dulled,” hence “obtuse”), though I can’t be confident not to have missed it. In its place (it seems) he employed the term fatuus, “insipid,” for bodies exhibiting neither acidic nor basic qualities (88; cf. 16). Winterl 1800, 2–3 (quote on 3). Winterl regularly used the word solvere and its related forms with the meaning of “neutralize,” only occasionally using neutralisatus (e.g. 69). Among his five categories of acids and bases was one for those acids (bzw. bases) which unite with other acids (bzw. bases). Arguing that their reaction differs from that of acids and bases with each other—whereby their tastes and power of changing the colors of pigments do not mutually blunt each other (“reciproce non obtundunt”), and the proportions in which they combine depend on a variety of external factors—he designated “this very different species of neutralization [solutionis species], opposite to that which joins bases and acids,” by the new term “synsomatia” (11–12). An entry in the index reads “Synsomatia how it differs from neutralization [Synsomatia quomodo a neutralisatione differat] (11),” implying the equivalence of solutio and neutralisatio (269). After concluding that “neither acids nor bases… have a trustworthy external characteristic [character],” he asserted that “the goal of this dissertation shall be [to determine] whether acids have a common internal characteristic (a hidden cause of a common attraction to bases)” (Winterl 1800, pp. 19–20). He occasionally spoke of the “principle of the causticity of bases” (73; see pp. 73–74 for further examples of his typical usage). Winterl 1800, 16 and passim. One is reminded here of Ørsted’s conception of opposing forces of indefinite strength existing undetected in bodies (see Caneva 1998, pp. 65–67). Winterl 1800, p. 17.

ØRSTED’S PRESENTATION OF OTHERS’

279

In Winterl’s view, the animating principles of acidity and basicity are bound (relegatus) to their substrate (substratum) with a force the strength of which determines the strength of the acid or base: the weaker the adhesion (adhæsio) of the acidic principle to its substrate, the stronger will be the taste of the acid and the more readily will the force of a base (vis Baseos) be able to separate it from its substrate.22 With regard to the process by which a stronger acid or base is added to a neutral salt (Sal neutrum) formed from a weaker acid or base, driving out the latter, Winterl asserted that there can be no doubt about “the actual passage [transitus] of the principle of acidity from the expelling to the expelled acid.”23 In other words, his principles were not themselves qualities or forces, but the substantial bearers of qualities capable of exerting forces. Yet in language foreshadowing his later treatment of the relationship between heat and the principles of acidity and basicity, Winterl fleetingly suggested a nonsubstantial, more forcelike understanding of the process of neutralization (although he did not use any such particular term for it in this context): “When a complete acid [Acidum integrum] and a complete base are united one to the other [inter se uniuntur], proportionate quantities of the principles animating each are consumed through reciprocal action, such that they are nowhere further to be found; nevertheless between them there always arises an increase in temperature, which is directly proportional to the reciprocal attraction of the acids and bases and inversely to the times in which the union of the acid with the base takes place.”24 His further elaboration of this relationship, however, left no doubt about the materiality of heat and its constituent principles. It is noteworthy that in these connections Winterl always used the neologism “Caloricum” then generally applied to the imponderable substance of heat (“caloric”), never the classical Latin “calor.” After passing in review a number of reactions involving temperature change, Winterl prepared the reader for his conclusion with an appeal to a basic principle of reasoning in chemistry: But whenever two things completely disappear and a third is found in a proportionate quantity in their place, it is safe to conclude that the third has arisen from the joining [nexus] of the two that disappear; one thus concludes that water consists of the substrates of vital air and inflammable air, because both of these ignited airs disappear, with water remaining.…

22

23

24

Winterl 1800, pp. 32–33, 91–92. An index entry states that “[t]he strength of an acid does not depend on the abundance of the principle of acidity (114) but on the looseness by which it adheres to its substrate (32)” (264). Winterl 1800, 94. Cf. 113–114: “[I]ndeed, acids can be very strongly augmented in their principle of acidity in such a way that this animating principle is bound very firmly to its substrate, the acid being nevertheless only very weak; and, on the contrary, their substrate can be altered in such a way that it very easily transfers the given principle of acidity to other bodies, which are thus converted into very strong acids.” He once spoke of “the mobility [Mobilitas] of the principle of acidity from one substrate to another” (133). Winterl 1800, p. 134. Winterl’s treatment of heat and its relationship to the principles of acidity and basicity occurs in the last chapter of the first prolusion, entitled “On the Nature of the Principles of Acidity and Basicity Removed from Their Substrates” (134–168). An index entry reads: “The principle of basicity common to all bases (133) is the cause of the reciprocal action of bases on acids (ibid.) the two of them conjoined constitute heat [ambo conjuncta componunt Caloricum] (134–140)” (267).

280

K. L. CANEVA

Anyone who would deny this inference would thereby destroy all of chemistry, in which whatever we know depends on the same method of reasoning. Heat, therefore, insofar as we are capable of understanding by means of the human mind, consists in the united principles of acidity and basicity: these two principles, from which alone all attraction in nature appears to depend, united one to the other, constitute a composite matter [materia composita] (heat [Caloricum]), whose unique action seems to be to penetrate all nonresisting places—which are the pores of bodies—by means of its absolute (unlimited) elasticity, and to repel from itself the parts of the penetrated body.25

Experiments—especially involving the compression and rarefaction of air— demonstrate that “heat can be expelled from the pores of bodies by compression like water from a sponge,” whence one can conclude that “heat cannot exist except in space empty of all matter, i.e. in pores.”26 On the purely phenomenological level, Winterl pointed out that the friction of different bodies can produce either electricity or heat, which he interpreted in terms of the principle that “friction increases all natural attractions.”27 He explained the production of heat by saying that friction alters the heat capacities of the bodies, which thereupon attract to themselves “the heat everywhere diffused and joined to no body.”28 Without apparent justification, he asserted that the constituent part of heat collected by a body in electrical connection with the earth is the principle of acidity, the part collected by an insulated body the principle of basicity, then further identified the principle of acidity with negative electricity.29 As far as I can tell, that’s as close as he got to a justification of his identification of the constituent parts of heat with the principles of acidity and basicity and thence with positive and negative electricity: he was, after all, writing before the advent of Volta’s pile and the wealth of electrochemical phenomena it unfolded.30 After a lengthy discussion of the mechanical generation of (static) electricity and its relationship with various chemical phenomena—the exposition of which I cannot wholly follow—he noted that the coming together of basic and acidic principles from the insulated and grounded conductors produces “glowing heat” in the form of a little flame, “which always falls downward from the superior to the inferior conductor…, which is an indication of the weight [gravitas] of the heat, and a new argument against fixed heat, which should—against experience—increase the body’s weight.”31 He declared that only light (lux) is without weight, and wondered “whether light is the common cause of magnetism and consequently of weight,” though without pursuing the matter any further.32 Such fleeting references were

25 26 27 28 29

30

31 32

Winterl 1800, pp. 139–140. Winterl 1800, p. 141. Winterl 1800, p. 149. Winterl 1800, p. 150. Winterl 1800, pp. 150–151, 154. He spoke without further ado about “the parts of heat separated by the electrical machine” (157). But see Winterl 1800, pp. 159. An index entry reads: “Electrical phenomena … commonly so called are not such (149, 159) but are a cause [causa] carried over into the insulated conductor (159, 160)” (263). Winterl 1800, 154; cf. 167–168 for additional chemical phenomena involving heat and electricity. Winterl 1800, pp. 158–159 (quote on 159); cf. 160.

ØRSTED’S PRESENTATION OF OTHERS’

281

to capture more of Ørsted’s attention than the more strictly chemical issues that made up the bulk of the book. It is important to recognize just how excited Ørsted was by the gathering prospects for a radical reformation of chemistry and physics following the discovery of Volta’s pile and the combined experimental findings and theoretical developments advanced by his close friend, Johann Wilhelm Ritter (1776–1810), and by the as yet unrecognized Winterl.33 For a few years it looked to him like Ritter and Winterl were showing the way towards a new science based on a new set of principles, principles in tune with the dynamical views that had attracted him to Kant’s philosophy of nature. Having seen at first hand the resistance to Winterl’s ideas among German scientists, Ørsted took upon himself the task of translating those ideas into German. The principal manifestation of that effort was his Materials for a Chemistry of the Nineteenth Century, whose “Erstes Stück” was tendered as the first installment of what was to be an extended exposition and testing of Winterl’s ideas. Expressing the desire “to make Winterl’s ideas more generally known,” Ørsted published what he hoped would be a “comprehensible presentation [faszliche Darstellung] of Winterl’s system.”34 That he succeeded is attested by an anonymous reviewer of Winterl’s works, who recorded that the Prolusiones had initially attracted little attention “until Dr. J. C. Ørsted made chemists more aware of it by means of a good extract from it” in his Materialien.35 In that I would judge him largely to have succeeded: in part by significantly paring down Winterl’s presentation of experimental examples, Ørsted offered the reader a more compact yet basically accurate account of the two principal aspects of Winterl’s work, namely his theory of acids and his claims for andronia. His vocabulary closely followed Winterl’s: Principium aciditatis became Princip der Acidität (or Säureprincip)—though for Principium basicitatis he preferred Princip der Alkalität (or Kausticitätsprincip)—while Principia animantia became belebende Principien; Caloricum was usually rendered as Wärme, sometimes as Wärmestoff or Wärmeprincip, with Glühhitze standing in for Caloricum lucidum; and aër vitalis survived as Lebensluft. Similarly, the reader of both works quickly recognizes the equivalence of obtundere and abstumpfen, of solutio and Neutralisation, of fatuus and fad, of Acidum integer and vollständige Säure. But those were not the aspects of Winterl’s work that spoke most directly to Ørsted’s own deepest interests, and it is precisely the biases in his presentation that are of most interest to us here.36 Winterl himself was apparently more bothered by what he took to be Ørsted’s distortion of his views than he was pleased by the increased exposure Ørsted had given them: responding to the above-cited

33

34

35 36

Cf. Jacobsen 2000, ch. 3, “Ritter’s Influence on Ørsted,” (47–64) and ch. 5, “Spirit and Unity— Ørsted’s Fascination by Winterl’s Chemistry” (87–121), the latter reworked as Jacobsen 2001. Ørsted 1803a, 143; cf. 1813, p. 17 = 1920a, 2, p. 175. Note that page references to the Materialien are not to the original printing, but to its republication in Ørsted’s Naturvidenskabelige Skrifter. It is, in fact, an extremely rare book. Anonymous 1806, col. 346. Jacobsen (2000, 100, 110, 120, 184; 2001, 197–198, 206, 212) has discussed the issue of Ørsted’s selective appropriation of Winterl’s concerns, especially those that were consonant with Ørsted’s Naturphilososphie.

282

K. L. CANEVA

anonymous reviewer’s claim that (in Winterl’s paraphrase) “Ørsted is supposed to have supplied a good extract from my Prolusions,” he countered that Ørsted had “mangled the context [er zerriß aber den Zusammenhang], outside of which individual claims can have no great weight.”37 Perhaps the most striking aspect of the preface to the Materialien is the fervor with which Ørsted presented Winterl’s system as an answer to a host of deficiencies in Lavoisier’s chemistry. Indeed, in an unpublished retrospect Ørsted recorded that “[a]lready in the year 1800 I began to feel the incompleteness of Lavoisier’s theory, which sees combustion only in terms of combinations with oxygen and leaves all the other phenomena if not entirely unexplained, then still only loosely connected with the whole.”38 He thus saw in Winterl’s work a viable route to a more adequate understanding of a broad range of interconnected phenomena. Hence what had originally been a controlled jab at Lavoisier’s theory of acids in Winterl’s preface became, in Ørsted’s hands, a broad assault against a host of weaknesses in a structure that, in Ørsted’s view, was fatally limited by its narrow focus on the relations of oxygen (here still “vital air”). He enumerated the questions he wanted answers to—questions, he said, that Lavoisier’s theory had no hope of ever answering: Why do acids and bases neutralize each other?—Why is electricity necessary for the combination of several kinds of gases with each other? Or in general: What is the relationship between the electrical phenomena and the chemical phenomena in association with which they have so often accidently been found, but which would be found much more often if one proceeded according to principles? Why does water always require an addition so that the calcination of a metal, etc., in it may occur? Does there follow from the antiphlogistic theory of combustion a theory of spontaneous combustion? Is there a thorough explanation of the phenomena of light and of changes in color in chemical processes? What is the common principle of metals? What that of alkalis and earths?39

37 38

39

Winterl 1806, 313. Ørsted 1812a, 223 (§229): “Allerede i Aaret 1800 begyndte jeg at føle Ufuldstændigheden af den Lavoisierske Lære, der ved Forbrændningen kun seer paa Forbindelsen med Surstoffet (Ilten), og lader alle andre Phænomener om ei ganske henstaae uforklarede, saa dog kun løseligen sammenknyttede med det Hele. Hvad der først vakte min Opmærksomhed var Voltas Støtte og Ritters galvaniske Arbeider før og med samme.” (I thank Anja Skaar Jacobsen for help with this passage.) He went on to name Winterl as “the other founder of the new theory [den anden Stifter af den nyere Lære].” Other evidence of his early dissatisfaction with Lavoisier’s chemistry comes from his autobiography: “Before he was 24 years old [which he became on 14 August 1801] Volta’s great discovery, Ritter’s brilliant works, and Winterl’s ambitious system [driftige Lærebygning] had already instilled in him the conviction that the antiphlogistic theory could not endure” (Ørsted 1828, 526). For a further discussion of Ørsted’s early identification of problems with antiphlogistic chemistry, see Jacobsen 2000, pp. 13–14, 90–96 or 2001, pp. 187–193. She makes the important point that, “[i]n contrast to Lavoisier, Winterl and Ørsted were more interested in reactivity than in composition” (2000, p. 91; cf. 2001, p. 211). Ørsted 1803a, pp. 139–149 (treating as continuous text items originally set off as separate paragraphs). In his travel diary of 1802 Ørsted referred to Ritter’s “discovery concerning the antiphlogistic delusion [Vildfarelse] regarding the composition of water” (Ørsted 1870, 1, p. 78). In a critique of eudiometry published three years later he revisited the issue: “If one accepts Ritter’s opinion about water, then there can be no doubt that oxygen gas [Surstofgas] owes its entire nature to electricity; if, on the contrary, one still stands with the antiphlogistians’ party, then one usually—and quite correctly—doesn’t want to hear anything about electricity in chemistry; but after the experiences of recent years one can scarcely any longer refuse to do so” (Ørsted 1805b, pp. 72–73 = 1920a, 1, p. 258; cf. 1805c, p. 387). Winterl joined Ritter in recognizing this deficiency in Lavoisier’s chemistry. The next year (1806), in following up on some of Winterl’s experiments, Ørsted again

ØRSTED’S PRESENTATION OF OTHERS’

283

To which he added in a footnote, giving voice to Winterl’s tacit (and quite traditional) assumption: “That which constitutes a class must have a common principle.” Although Ørsted did not make a big issue of it, he clearly ignored Winterl’s distancing himself from dynamism by imputing a dynamical, anti-atomistic slant to his work. Winterl, he said, was concerned not only with the connections between individual experiments, but with “the connection of the individual to the whole… (taking it not atomistically but dynamically).”40 Without naming names, but in response to Ørsted’s and others’ attempts to enlist Winterl into the ranks of the dynamists, his student, Johann Constantin Schuster (1777–1838), insisted on distancing his teacher from Naturphilosophie, adding “[t]hat some Naturphilosophen adopted his system proves nothing for or against it.”41 The year before, in 1806, an anonymous reviewer of Winterl’s works insisted that “[t]he charge that his views owed their origin to the new, in part so disparaged Naturphilosophie is completely baseless.”42 Such misperception, of which Ørsted was a prime source, was clearly widespread. Although, as noted, Ørsted was careful to adapt his language in most regards to Winterl’s, he deviated from it in a few subtle ways that suggested a closer connection between Winterl’s system and the favored conceits of dynamic Naturphilosophie. When Ørsted wrote that “die Acidität und Alkalität heben einander auf,” that “die Alkalität und Acidität sind einander entgegengsetzt,” and that we would not know “dasz die verschiedenen Elektricitäten, Magnetismen, u.s.w., entgegengesetzt sind, wenn sie sich nicht gegenseitig aufhöben,” he used terms that in one case had no similarly nuanced Latin correlate (aufheben), in another had been used only sparingly (taking entgegengesetzt as equivalent to oppositus), but which were both indispensable and characteristic components of the language of contemporary dynamism.43 Certainly Winterl had never spoken of the decomposition of an acid into “Gegensätze,” nor of one of the products as being “in einem entgegengesetzten Zustand” vis-à-vis the other.44 When Ørsted spoke of electricity and magnetism

40 41

42 43 44

revisited the issue: “I am convinced that the whole theory of heat, such as the antiphlogistians have established it, must undergo a great revolution, the path to which Ritter and Winterl, each in his own way, have [already] prepared” (Ørsted 1806b, p. 285 = 1920a, 1, p. 283). Left unexplained by antiphlogistic chemists was the action of the galvanic pile on water, whereby both heat and air (Luft) are produced. Ørsted 1803a, p. 141; cf. Jacobsen 2000, 99 or 2001, pp. 196–197. Schuster 1807, 1, p. 12. As Jacobsen (2001, 212) observed, “Winterl’s system was for Ørsted a crucial example of applied Naturphilosophie, although it seems clear that Winterl’s chemistry was, at least by Winterl himself and Schuster, only associated with Schelling’s Naturphilosophie after Ørsted’s and Ritter’s engagement in it.” Schuster presented Winterl as the person who finally succeeded in establishing dualism—based on the Gegensatz between acid and alkali—on a broad base in science (1, pp. 94–95). In a later work, in which he adumbrated a radically new ontology—his “ousial” hypothesis—based on the assumption of one qualityless material substance (Stoff) and six qualitybearing “geistige Substanzen,” Winterl remarked that his view should be called “the heptalic (not dualistic, as Schuster called it)” (Winterl 1808, p. 3–5; cf. 269–270). Winterl apparently resisted being identified with any generally recognized schools of thought. Anonymous 1806, col. 347. Ørsted 1803a, p. 147. Ørsted 1803a, p. 156. Ørsted’s identification of thelyca as “der Gegensatz der Andronia” went beyond Winterl’s explicit characterization (172).

284

K. L. CANEVA

as “Kräfte,” and asserted that the material substances (Materien) to which they are bound do not need to be bound together chemically “damit sich die Kräfte aufheben könnten,” and when he referred to the erstwhile animating principles of acids and bases as “ihre belebende Kräfte,” he was using a concept of force either not systematically deployed in the Prolusiones, or, with regard to the notion of an Aufhebung der Kräfte and the glossing of a principle as a force, wholly absent from it.45 When he wrote Das [belebende Princip] der einen Klasse ist dem der andern entgegengesetzt. Wenn Säuren und Basen mit einander in Conflict kommen, heben sich daher ihre Charaktere wechselseitig auf.

—where the italics are his—he was not so much distorting Winterl’s ideas as expressing them in language suggestive of a framework of thought that was not Winterl’s.46 For the rest, Winterl never employed the concept of “conflict” that was to remain so dear to Ørsted for so long. It is worth noting that in later years Ørsted, with apparently increasing consistency, associated Winterl’s theory with the concept of force, a misrepresentation that betrays the route by which he himself had assimilated Winterl’s ideas as it surely has contributed to others’ misunderstanding of work few people bothered to read directly. Thus in his unpublished “Theory of Force [Kraftlære]” of 1812 Ørsted acknowledged that “[Winterl’s] theory of heat—that it is produced by the unification of the two opposite forces—I made my own.”47 The next year, in his Researches on the Identity of the Chemical and Electrical Forces, he testified for the umpteenth time to the revolutionary significance of the work of Volta, Ritter, and Winterl, now in explicitly “dynamical” terms: One has just seen that the first idea of the new system, which we call the dynamical system, came much before the discovery of galvanism, although in a very imperfect state. The development given to it by the works of Ritter and Winterl already date from the period preceding Volta’s pile. The experiments of the former on the simple galvanic chain had already led him to conceive the idea of an electrochemical theory; and the latter had inferred from chemical and ordinary electrical facts the identity of the forces that produce them, while the physicist from Pavia had given to chemistry this instrument, pregnant with new discoveries, which has transformed into certainties the fundamental idea of the dynamical system.48

In 1816 he introduced an exposition of his theory of light with a bit of history: “As a consequence of the discoveries with which the last twenty years’ endeavors have enriched science, one will no longer deny that the forces manifest in electrical actions are universal forces of nature, not different from the chemical forces. The author assumes now with Winterl that the unification of these two forces produces both heat and light.”49 And in 1822, in an overview of the history of 45

46 47

48 49

Ørsted 1803a, 147. Cf. 159, where he spoke of a metal being “mit der entgegengesetzten Kraft begabt.” His employment of “Kraft” with regard to the strength of a given body’s acidity or basicity is in line with Winterl’s occasional usage (159–161). Ørsted 1803a, p. 156. Ørsted 1812a, p. 223 (§229): “Hans Lære om Varmen, at den frembringes ved Foreningen mellem de to modsatte Kræfter, har jeg gjort til min”; translated in Jacobsen 2000, p. 184. Ørsted 1813, 12–13 = 1920a, 2, p. 173. Ørsted 1816, p. 12 = 1823a, p. xvi = 1920a, 2, p. 433.

ØRSTED’S PRESENTATION OF OTHERS’

285

chemistry, he wrote that “[i]n the same year that Volta’s pile was discovered, the Hungarian investigator, Winterl, made his appearance with a work in which he demonstrated the unity of the electrical and chemical forces in a wholly different way.”50 Such had never been Winterl’s way of speaking. Toward the end of the Materialien proper Ørsted came to consider Winterl’s views on the relationship among heat, the principles of acidity and basicity, and the two electricities, positive and negative. For the most part his presentation closely followed the original (except for his reversing Winterl’s association of positive and negative electricity with the principles of basicity and acidity, respectively): Just as heat [Wärme] can be composed out of its two principles, so too can it again be decomposed into them. A means to this is friction, which collects the heat in the rubbed bodies, and, when they are of heterogeneous nature like glass and metal, decomposes them such that the former attracts the principle of acidity, the latter that of alkalinity. When the metal stands in conductive connection with the earth, it receives continuously from it the alkalinity principle, in that the glass attracts the acid principle. The former becomes negative, the latter positive. The sour taste that the positive conductor excites, and the opposite from the negative, is sufficiently well known.51

So far so good. The next sentence, however, shifts the terms of the discussion out of Winterl’s language into the favored conceits of Naturphilosophie: “The combination of the two electricities with each other is the transition of the principles of heat from difference to indifference.” Ørsted’s more extensive discussion of these issues came in a sort of appendix to the main body of the book headed “Letter to a Friend on Winterl’s Prolusiones,” originally intended for separate publication in Ludwig Wilhelm Gilbert’s (1769–1824) Annalen der Physik.52 To judge only from it, one might think that Winterl’s book itself had been essentially, indeed exclusively, concerned with just those issues of dearest concern to Ørsted. Although the exposition is basically faithful to Winterl, it is couched very explicitly in terms of “Hauptsätze” ostensibly representing Winterl’s views pertaining to electricity and heat, although Winterl himself had never spoken of laws, theorems, or principles (in this sense) at all, let alone principal laws, much less principal laws relating to electricity and heat. The first Hauptsatz so identified stated that “[t]he principles of electricity are also the causes of acidity and alkalinity.”53 The second of the Hauptsätze, Ørsted 50 51 52

53

Ørsted 1822, cited from 1920a, 3, p. 302. Ørsted 1803a, p. 199. The friend was identified as Ludvig Manthey in Ørsted 1813, p. 17 = 1920a, 2, p. 175. As recounted by Meyer (1920, xxviii, no further source given), “[Ørsted] sent [the “Letter”] to Ritter in order that he might get it published in Gilbert’s ‘Annalen der Physik,’ but Ritter thought it better that the letter should be published as the last chapter of the book, as he himself purposed [sic] to write something about it that might take the place of a review in a paper he was writing for Gilbert.” Her source was Ørsted’s then-unpublished letter of 28 October 1802 to Ritter (Ørsted 1920b, 2, pp. 28–29). Cf. Jacobsen, 2000, 90, who cites both Meyer and Ørsted 1920b, 2, 29. According to Jacobsen (2000, pp. 106–107 = 2001, p. 203), the anonymous reviewer of Ørsted’s Materialien in Trommsdorff’s Allgemeine chemische Bibliothek des neunzehnten Jahrhunderts (4, 1804, pp. 127–141) expressed the opinion that “Winterl’s system appeared much more appealing from Ørsted’s short “Letter to a Friend” than in its complete extensiveness,” testifying to the effectiveness of Ørsted’s recasting. On this review and the significance of Ørsted’s “Letter” see also Meyer 1920, p. xxviii. Ørsted 1803a, p. 206. In his book of 1812 Ørsted stated that “Winterl first expressed the happy idea that caloric [Wäremestoff] is produced through the unification of the principles of alkalinity and

286

K. L. CANEVA

asserted, deserved to be placed at the head of the whole system: “The principle of heat is composed of the principle of acidity and that of alkalinity.”54 Winterl had never so identified or so elevated it. Again we find ourselves in an un-Winterlian linguistic world of entgegengesetzt, aufheben, and Indifferenz, now even of “the unity of all forces.”55 The paragraph in which the latter phrase occurred might possibly have corresponded to Winterl’s private beliefs, but it most assuredly did not reflect the tone of his Prolusiones: The constituent principles of heat that play a role in alkalis and acids, in electricity, and in light are also the principles of magnetism, and in such a fashion we would then have the unity of all forces which, intimately interconnected, rule the entire structure of the world, and the physical knowledge so far attained thus becomes a unified physics formed out of one piece…, for do not friction and percussion produce both heat and electricity, and do not dynamics and mechanics thereby become completely and intimately interconnected?…Our physics will thus no longer be a collection of fragments—on motion, on heat, on air, on light, on electricity, on magnetism, and on who knows what more—rather we shall embrace the whole world with one system.56

Such was the vision that ever hovered out of reach before Ørsted’s eyes. Ørsted seized the opportunity for another extended synopsis of Winterl’s work in an essay devoted to an “Overview of the Most Recent Progress in Physics,” published anonymously the same year (1803) in the second issue of Friedrich Schlegel’s journal, Europa.57 As far as this essay was concerned, that recent progress consisted almost entirely in the work of Ritter and Winterl: those were the men whose discoveries and ideas were going to transform physics. Ørsted began his ten-page précis of the Prolusiones’ “Hauptsätze”—only the first of which was explicitly so tagged—with a subtly transformed version of the two he’d announced in the Materialien. We are now told that “[t]he cause of heat is the product of the two electrical principles,” not that the “principle” of heat is so “composed,” implicitly leaving open the question as to the nature of heat.58 Going beyond Winterl, he

54 55 56

57

58

acidity” (1812b, p. 153 = 1920a, 2, p. 105). In the book’s French translation this became “Winterl was the first who had the happy idea of considering caloric [calorique] as composed of the two principles of electricity” (1813, p. 144). Ørsted 1803a, p. 208. Ørsted 1803a, p. 209 (“die Einheit aller Kräfte”). Ørsted 1803a, p. 209–210. The entire passage is translated in Stauffer 1957, p. 38. My translation is freer than I usually like. The second ellipsis contained an unspecific parenthetical reference to pertinent experiments of Ritter’s. The essay appears to have been written considerably after the Materialien but published—or delivered to the publisher—before it, as its anonymous author referred the reader to Ørsted’s book, “which, according to the announcement, will soon be coming out” (Ørsted1803c, p. 45 = 1920a, 1, p. 129). Schlegel was then living in Paris, where he had contact with Ørsted; they had met two years earlier while Ørsted was passing through Jena (Ørsted, travel diary, entries for 22 September 1801 and, 18 May 1802, in Ørsted 1870, 1, pp. 26, 90, 138). Ritter repeatedly asked Ørsted to greet the Schlegels (letters of 28 October 1802, 20 May 1803, 15 June 1803, and 20 June 1803, in Ørsted 1920b, 2, pp. 31, 40, 43, 45). A few years earlier, Schlegel had hoped to see from Ritter’s hand “eine historische Uebersicht der Chemie” (Friedrich Schlegel, letter to August Wilhelm Schlegel of 4 September 1800, in Schlegel 1890, p. 438). Ørsted’s essay was perhaps a stand-in for the piece Ritter never wrote. Ørsted 1803c, p. 35 = 1920a, 1, p. 123. In noting that Ritter had assigned the principal role to light where Winterl had placed heat, Ørsted suggested “that light and heat are the results of the same principles” (41 bzw. 127). In noting Winterl’s argument for the materiality and ponderability of the

ØRSTED’S PRESENTATION OF OTHERS’

287

now allowed himself to draw evidence from phenomena associated with Volta’s “elektrische Säule.” He was quietly moving from exposition to elaboration. In an 1806 paper on “The Series of Acids and Bases” Ørsted again appealed to the findings and theories of Ritter and Winterl in an exploratory attempt to define acids and bases and to arrange them in a single series reflecting their relative strengths. In addition to the Prolusiones he now also cited Winterl’s Exposition of the Four Components of Inorganic Nature (1804) and Schuster’s as yet unpublished System of Dualistic Chemistry of Professor Jakob Joseph Winterl (1807). Asserting that the business of the true Naturforscher is the search for the “inner unity” of similar phenomena, Ørsted credited Winterl (“this profound investigator”) with the recognition that no definition had yet captured that unity with regard to acids and bases, chemists having been limited to noting the differing boundaries implied by the use of different manifestations of their action.59 His intention was to use the property of acids and bases to neutralize (aufheben) each other’s action, but ran into the problem that some of the weakest bases require the greatest quantities of acid for their saturation (Sättigung), and in the event he had no way to distinguish between what he recognized must be a distinction between the quantity and intensity of the acidity.60 As usual, he kept his ontological options open: “To express ourselves clearly, we would say to the supporters of the customary view that the principle of acidity or basicity is more bound in the one substance than in the other; to the friends of the dynamical view, on the other hand, we would declare that the same force either concentrates itself more towards the inside or strives to disperse itself towards the outside, according to whether the internal conductivity is greater or smaller.”61 Noting that “Winterl has…proven—and galvanic experiments have confirmed it—that the opposite electricities are the principles of acidity and basicity, and that their unification produces heat,” Ørsted broached the prospect that measurement of the heat produced in acid-base reactions might provide a means of gauging their relative strength, but he made no attempt actually to measure or to quantify that heat.62 In the end he believed that what constitutes an acid or a base cannot simply be determined by the presence or not of the principles of acidity and basicity, identified with negative and positive electricity, respectively—for they are present in all bodies—but on the manner of their presence (“die Art ihres Daseyns”), on their different “forms” as revealed, for example, in the phenomena of light, heat, electricity, magnetism, “etc.”63 Having

59 60 61

62 63

principle of heat, Ørsted cautioned against any too-hasty conclusion: “This materiality cannot be rejected by any mere philosophy, for even if the latter had proven an immaterial principle a priori, it would nevertheless not thereby be settled that precisely the phenomena of heat, light, etc., derived therefrom, rather this would still have to be left to experimental investigation. One should not take this remark as a defense of the materiality of the heat principle, but only as a necessary remark on the conditions of this contentious question” (41–42 bzw. 127). Ørsted 1806c, p. 513 = 1920a, 1, p. 292 (“innere Einheit”; “dieser tiefsinnige Forscher”). Ørsted 1806c, p. 519 = 1920a, 1, p. 297. Ørsted 1806c, pp. 520–521 = 1920a, 1, p. 297. A footnote to this passage refers the reader to Schuster 1807, 2, pp. 66–79. Ørsted 1806c, p. 530 = 1920a, 1, p. 303. Ørsted 1806c, pp. 538–539 (quotes on 539 and 538, respectively) = 1920a, 1, 309.

288

K. L. CANEVA

been mightily inspired by his interpretation of Winterl’s ideas, in the end Ørsted never really found a way to make good on their early promise. Sympathetic readers of Ørsted’s enthusiastic treatment of Winterl’s chemistry would likely have come away with the idea that it constituted a major theoretical challenge to the reigning antiphlogistic system, and that it offered, at the least, insightful suggestions of how to proceed in the construction of an adequate and more inclusive replacement. Unless they were among the apparently rare few who had read Winterl’s original works, they would probably not have realized that the Winterl they confronted had been significantly modified by having passed through Ørsted’s dynamically refracting lens. Thus Irish chemist Richard Chenevix (1774–1830), a well-known scourge of all things naturphilosophisch, in opening his widely published critique of Ørsted’s Materialien by stating that it was nothing but a summary of Winterl’s Prolusiones, assimilated Winterl’s original to Ørsted’s reworking.64 That assimilation included a foregrounding of Ørsted’s critique of Lavoisian chemistry, the representation of the principles of acidity and alkalinity as “opposite forces” capable of destroying each other, and, most significantly, an extended and explicit identification of Winterl’s chemistry as an expression of the “sect”—Ausgeburt in the German translation–of philosophy that had for some time “ravaged” parts of Germany and whose leading ideas Chenevix subjected to a biting parody.65 A passage from Hermbstädt’s review of Ørsted’s Materialien reveals the extent to which Ørsted had transformed Winterl’s system into a species of Naturphilosophie: “An andronia [and] a thelycke, a principle of acidity and [a principle of] alkalinity that neutralize each other or bring each other to [a state of] indifference, afford the sought-for duplicities, the conflicts and indifferences, that comprise the whole game of Schellingianism.”66 No such reading would have been possible of Winterl’s own words. To a very considerable extent, the Winterl who has come down to us is the Ørstedized version closely associated with the dynamical Naturphilosophie Winterl himself stood apart from.

3. ØRSTED’S PRESENTATION OF RITTER’S EXPERIMENTS ON THE CHEMICAL ACTION OF LIGHT Prompted by William Herschel’s (1738–1822) detection in 1800 of calorific rays beyond the red end of the solar spectrum and his own conviction that polarities underlie the principal phenomena of nature, Ritter thought to see if he could detect invisible solar rays also at the other end of the spectrum. The first published

64 65

66

Chenevix 1803–1804, 5, (1803), 241 = 1804, 173, 1805, 422. Chenevix 1803–1804, 5 (1803), pp. 242, 243, and 6 (1804), pp. 7–12 1804, pp. 173–174, 175, and 191–196, 1805, pp. 422–423, 424, and 442–449 (which quotes several of the purplest passages in the original French in footnotes to Gilbert’s rather free translation). Hermbstädt 1804, p. 469, quoted in Meyer 1920, p. xxvii. Hermbstädt went on to characterize the spirit of Winterl’s theory as “alchemical” insofar as it sought to explain the common properties of substances in terms of their possession of peculiar substances (“eigene Stoffe”) (Hermbstädt 1804, p. 469).

ØRSTED’S PRESENTATION OF OTHERS’

289

account of his findings appeared as a note in the Erlangen Litteratur-Zeitung of April 18, 1801 with the revealing title, “Chemical Polarity in Light,” in which he summarized his findings under the following three heads: 1. There are rays in sunlight that do not illuminate and of which one part is refracted more strongly, the other more weakly, than all those that illuminate. 2. Sunlight in the undivided state is a neutralization of the two ultimate determinants of all chemical activity, oxygeneity and deoxygeneity (= hydrogeneity). 3. By means of the prism the two diverge like poles. The red side of the spectrum and that which borders on it externally become the side of oxygeneity, the violet side, on the contrary, and that which borders on it become the side of hydrogeneity. The maxima of both fall outside the visible spectrum; their indifference, however, [falls] inside it in the region of green.67 What most excited Ritter was the evidence his discovery provided for the thoroughgoing polarity of all the activities of nature and for their complex interconnections: “It will be the result of a larger factual investigation to exhibit the polarity of chemistry, electricity, galvanism, magnetism, heat, etc., in accordance with their principles as one and the same in all.”68 The fuller elaboration of his discoveries that he presented that spring to the Jena Naturforschende Gesellschaft (although not published till 1806, in a three-volume collection of Ritter’s papers) abundantly exhibited the naturphilosophisch extent to which Ritter’s work represented to him much more than the discovery of the still unnamed ultraviolet light. On the basis of his new findings one could now conclude that [t]he entire spectrum of the prism appears in a new dignity as chemical. In white light the two forces, the one that determines oxidation and the one that determines deoxidation, are in a state of mutual combination. The prism divides them; in in-each-other becomes alongsideeach-other; they separate like + and − ; the entire spectrum breaks down into two parts, of which one, B, becomes the sphere of oxidation, the other, A, that of deoxidation.69 In his final paper on the subject, published in the December 1802 issue of the Annalen der Physik, Ritter recapitulated his discoveries in language that repeated his earlier characterizations as it emphasized the fundamental distinctiveness of the invisible chemical and thermal rays he and Herschel had discovered vis-à-vis the light of the visible spectrum.70 Unless he had happened to catch the one-paragraph excerpt from a letter of Ritter’s in the April 1801 issue of the Annalen der Physik, it is unlikely that Ørsted knew of Ritter’s work on light before making his acquaintance during his

67

68 69 70

Ritter 1801a, col. 123. In a personal communication, Andreas Kleinert informed me that the obscure words “immer dasselbe” as printed in the original (in the last phrase quoted here) are a misprint for the (grammatically irregular) “inner dasselbe,” as he confirmed by examining the corrected copy of the text in the Erlangen University Library. This section is based on Caneva 2001, 11–17. Ritter 1801a, col. 123. Ritter, 1806b, p. 92. The accompanying diagram is reproduced in Caneva 1997, p. 46. Ritter 1802b, pp. 409–411, 413 = 1806a, 2, pp. 355–357, 359. According to Kleinert (1984, pp. 295 and 298, n. 24) this issue probably appeared no earlier than April 1803.

290

K. L. CANEVA

travel abroad from August 1801 to January 1804.71 From their first encounter, in September 1801, Ørsted recorded only that Ritter had shown him “his most remarkable experiments.”72 It was only after a longer return visit to Jena—from 13 August to 4 September 1802—that Ørsted indicated any knowledge of Ritter’s spectral investigations: “[H]is experiments on light, which are so infinitely important, have not found a single person who has tried to repeat them.”73 In a pattern to be repeated, Ritter’s close friend took on the task of explaining his friend’s work to the scientific public. In what may have been his first such endeavor, Ørsted, in Paris from roughly the beginning of December 1802 until the end of October 1803, composed a note that was published in the issue of the Société Philomatique’s Bulletin des Sciences for Germinal, an XI (= March/April 1803) with the identification “Note communiquée par M. Vicktred, docteur à l’université de Copenhague” after the title, “Experiments on the Invisible Rays of the Solar Spectrum.”74 Ørsted’s was clearly not a name yet readily recognized by his Parisian colleagues. This appears to have been the first announcement of this work of Ritter’s to appear in French. Presumably tailoring his account to what he anticipated would be the rather positivistic leanings of French scientists, Ørsted’s note began with the simple assertion that “[t]hese researches form the sequel to those by which Herschell [sic] recognized the existence of calorific rays outside of the solar spectrum.”75 He recounted Ritter’s experiments with the darkening of muriate of silver beyond the violet end of the spectrum and with the whitening of already slightly darkened muriate of silver in the invisible rays beyond the red end of the spectrum, the latter being, he said, thereby “deoxygenated [déoxigéné].”76 From these and several other reported experiments Ritter concluded “that there exist, outside of the spectrum and at its two extremities, invisible rays that possess the property of favoring oxygenation and deoxygenation.”77 The immediately following conclusion to this short note modestly extended the range of the implied polarities: “The same physicist has found remarkable relationships between these effects and those of metallic electricity. According to him, when the eye is in contact for some time with the negative conductor of a pile it sees objects red; in contact with the positive conductor, it sees all objects blue, from which there 71 72 73 74

75 76 77

See Ritter 1801b. Ørsted, travel diary, entry for 19 September 1801, in Ørsted 1870, 1, p. 25. Ørsted, travel diary, in Ørsted 1870, 1, p. 79; quoted in K. Nielsen 1989–1991, 20, p. 146. Ørsted 1803b. My identification of the author is based on the following considerations: (1) Neither I nor my agents in Copenhagen—I thank Thomas Söderqvist in this regard—have been able to identify anyone with such a name, nor does it appear likely to have been the name of any Dane. (2) It is highly unlikely that there was a second “docteur à l’université de Copenhague” who, in 1803, was privy to Ritter’s work. (3) Several of Ørsted’s contributions to the Journal de Physique that year—e.g. Ørsted 1803f and 1803i—identified him as “docteur à l’université de Copenhague.” (4) It is extremely common in French sources of the day for foreign names to be garbled, and Ørsted would then unlikely have been much known in Paris. (5) If one looks at examples of Ørsted’s signature, it is easy to see how the initial Ø might be misread as a V, and the overall visual impression of “Ørsted” is quite close to “Vicktred”—especially if one posits an unrecorded intermediary form “Victed,” which became further distorted on its way into print. Compare the signature reproduced under an engraved portrait of Ørsted from 1803 in Ørsted 1920a, 1, frontispiece. Ørsted 1803b, p. 197, 1803d, p. 255. Ørsted 1803b, p. 197, 1803d, p. 255 (“disoxigenated”). Ørsted 1803b, p. 198, 1803d, p. 255.

ØRSTED’S PRESENTATION OF OTHERS’

291

would seem to result an analogy between the action of negative electricity and red light, and between positive [electricity] and violet light.”78 That note may have been the occasion for Ørsted’s invitation to deliver a longer report to the Société Philomatique on 18 March 1803, published some months later in the Journal de Physique in what was the most substantial account of Ritter’s work to appear in French, none of Ritter’s own relevant works ever having been translated.79 As he would be for Ritter’s galvanic experiments, so here, too, Ørsted was Ritter’s French voice.80 Here again Ritter’s work—which Ørsted said was little known, even in Germany—was presented in connection with Herschel’s, described in straightforward terms, and interpreted in the first instance in terms of the contrast between oxygenation at the red end and deoxygenation at the violet end of the solar spectrum. Again a paragraph noted the effects on the eye’s perception of color of stimulation by positive and negative electricity, but expanded somewhat on the analogical connections: “The chemical action of positive electricity is also the same as that of red light; that is, they both favor oxygenation. Negative electricity and the violet ray preserve the same analogy in that they both favor deoxygenation—which experiments with the Voltaic pile have sufficiently made known.”81 The tone of Ørsted’s exposition of Ritter’s work on light was dramatically different in an article composed for a German audience, the one he published in Schlegel’s journal, Europa, that also contained an extensive exposition of Winterl’s chemistry. Ritter’s work provided him his first example of recent progress in physics—indeed, the only person’s he discussed at any length other than Winterl’s. His first few sentences were basically descriptive even as they underscored the phenomena of special importance to Ritter, the chemical polarity of light: Violet light is among all light rays the most deoxidizing; this the experiments of Scheele had taught us. Red light is accompanied by the greatest heating; Herschel showed us that and at the same time proved that alongside red light there are invisible rays that most strongly possess the heating capacity. But these discoveries stood there as yet very isolated, without connection with the rest of the phenomena, when Ritter discovered that there are invisible rays on both sides of the spectrum; that the ones on the violet side bring about deoxidation, the ones on the red side oxidation, and that the closer the rays are to the red, the more they promote oxidation; but the closer they are to the violet, the more they promote deoxidation.82

From there, Ørsted went on to connect these phenomena with others Ritter had discovered with regard to the effect of galvanism on the eye: when the optic nerve is rendered positive, objects appear red and larger than normal; when rendered

78 79

80

81

82

Ørsted 1803b, p. 198, 1803d, p. 256. Ørsted reported that “[t]his notice was read at the Société Philomatique several months ago” (Ørsted 1803g, p. 410 = 1920a, 1, p. 247). It was published in the issue of Frimaire, an XII (= November/ December 1803). In his travel diary Ørsted wrote with regard to 18 March 1803: “In the evening at the Philomatic Society, where a report I supplied over Ritter’s experiments with light was read, not without approval” (1870, 1, p. 125). Ørsted’s several reports on Ritter’s work almost invariably appear listed (as in the Royal Society Catalogue of Scientific Papers) under Ritter’s name. Ørsted 1803g, p. 410 (= 1920a, 1, pp. 246–247), 1804a, p. 215 (which clarifies the garbled punctuation of the original). Ørsted 1803c, pp. 20–21 = 1920a, 1, p. 113.

292

K. L. CANEVA

negative, objects appear blue and smaller than normal. Still other phenomena established a network of relationships among a variety of polarities: oxidation vs. deoxidation, positive vs. negative pole, acids vs. bases, expansion vs. contraction, Chladni’s discovery that the pitch of a flute is higher in hydrogen than in oxygen, and the contrasting effects of positive and negative galvanism on the tongue and the pulse. Again it was the polar interconnections among a diversity of phenomena that was most significant about Ritter’s work, not simply the discovery of invisible chemically active rays beyond the violet end of the solar spectrum.83 In a later work the significance of these phenomena was located in their demonstration of the “chemical partitioning [Vertheilung]” of light into more oxidizing and more reducing parts attendant upon its passage through a prism, of what Ørsted also referred to as “the partitioning of forces [Kraftvertheilung] in light.”84 The light under which Ritter’s work was seen was soon colored by the contemporaneous work of William Hyde Wollaston (1766–1828). Like Ritter inspired by Herschel’s discovery in 1800 of calorific rays less refrangible than red light, and guided by Scheele’s experiments with muriate of silver, Wollaston reported in June 1802 that “on the other [side of the solar spectrum] I have myself observed, (and the same remark has been made by Mr. Ritter,) that there are likewise invisible rays of another kind, that are more refracted than the violet. It is by their chemical effects alone that the existence of these can be discovered.”85 Finding that the blackening of the silver chloride extended far beyond the violet end of the spectrum and that “by narrowing the pencil of light received by the prism, the discoloration may be made to fall almost entirely beyond the violet,” Wollaston concluded “that this and other effects usually attributed to light, are not in fact owing to any of the rays usually perceived, but to invisible rays that accompany them; and that, if we include two kinds that are invisible, we may distinguish, upon the whole, six species of rays into which a sun-beam is divisible by refraction.”86 Although there were to be a few dissenters, Wollaston’s insistence on the essential distinctiveness of chemically active and luminous (visible) rays was generally subscribed to until the 1840s. Prompted by the publication of Ørsted’s report on Ritter’s “Experiments on Light” in the July 1804 issue of William Nicholson’s (1753–1815) Journal of Natural Philosophy, Wollaston urged “caution against the theory implied by the term

83

84 85 86

In an “Invitation to Physical and Chemical Lectures” of 1804, among other examples of how physics had shown in experience aspects of the world that were otherwise inaccessible, Ørsted noted that “Ritter showed us chemical actions and the modifications of light as changes in polarity, and found the transition from these to electricity and thence to magnetism” (1804c, 620 = 1920a, 3, p. 78). In other works of the period, Ørsted exploited Ritter’s findings as support for his attempt to understand light and colors in their connection with heat, electricity, and chemical activities, in particular in terms of “the connection we find throughout nature between force and form”; see Ørsted 1805a, 20–21 (= 1920a, 3, pp. 104–105), quote on 21 (bzw. 105) and 1805d, pp. 257–258 (= 1920a, 3, pp. 113–114), both translated in Caneva 1998, 61 and 69. Privately, Ritter contemplated an elaborate schematization of the (now tripartite!) relations among light, chemical action, electricity, magnetism, and heat (letter of 23 May 1803 to Ørsted, in Ørsted 1920b, 2, p. 39). Ørsted 1812b pp. 216 and 217 = 1920a, 2, p. 133. Wollaston 1802, p. 379 = 1803, p. 100, in a footnote. Wollaston 1802, p. 380 = 1803, p. 100.

ØRSTED’S PRESENTATION OF OTHERS’

293

‘disoxidating’ as applied to those rays.”87 Ørsted’s report had not in fact characterized the new rays as (adjectivally) disoxidating, but it did assert that “[n]egative electricity and the violet ray possess the same analogy, both promoting disoxigenation,” as it likewise claimed an analogy between red light and “oxigenation.”88 Against that specific designation, Wollaston cited experiments “which prove that the same rays, which cause the emission of oxygen by muriate of silver, occasion its absorption by the resin usually called gum guaiacum.”89 He hence favored qualifying those rays more generally simply as “chemical,” recalling here the tacit usage he had employed in his first communication on the subject.90 His suggestion was to be widely followed, giving terminological reinforcement to the sense that these rays were essentially and qualitatively different from visible light. In other words, despite Ritter’s richly ramified `understanding of the significance of his experiments with light, electricity, chemical action, and physiological response, an understanding strongly seconded by Ørsted in his German and Danish publications, in the near term—lasting until the 1840s and 1850s—Ritter’s work came to be seen as nothing more than the discovery of chemically active rays beyond the violet end of the solar spectrum. Even the characterization of those rays as deoxidizing—as Ørsted emphasized in his expositions of Ritter’s work to French audiences as the safest experimentally demonstrable minimum that might be claimed before an audience deemed undisposed to applaud bold speculative analogizing—did not survive. In the event, neither Ørsted’s sometime seconding of Ritter’s conception of the wider—in particular analogical—connections of his work on light nor his more modest characterization of them as deoxidizing was sufficient to define the terms under which that work was generally received. Ørsted may well have contributed to peoples’s knowledge of Ritter’s work on light, but he did not succeed in getting it accepted on Ritter’s terms.

4. ØRSTED’S PRESENTATION OF RITTER’S STORAGE COLUMN In addition to publicizing Ritter’s work on the chemical action of light, while in Paris during most of 1803 Ørsted also put a great deal of effort into presenting his friend’s galvanic and magnetic researches, in particular Ritter’s work relating to his Ladungssäule, the pile à charger (or pile secondaire) of Ørsted’s French publications.91 Both men believed that the discovery of that new device, and the further discovery it ostensibly made possible of the earth’s electrical poles, would

87 88 89

90 91

Wollaston 1804, p. 293 = 1811, p. 291. Ørsted 1804a, p. 215. Wollaston 1804, p. 293 = 1811, p. 292. Notwithstanding this counterinstance to Ritter’s characterization, Wollaston insisted that “[t]he experiments on guaiacum nevertheless prove distinctly, that the powers of the two extremities of the spectrum are not only different, but opposite in their chemical effects” (295 bzw. 296). Wollaston 1804, p. 293 = 1811, p. 292. For Ørsted’s presentation of Ritter’s galvanic and magnetic researches on topics other than his storage column see Ørsted 1803h and 1803i.

294

K. L. CANEVA

win for Ritter the Institute’s 3,000-franc galvanism prize. Ørsted’s role will best be appreciated after a review of the chronology of events. A month or two after Ørsted had left Jena early in September 1802—after three weeks of intensive collaboration with Ritter—Ritter read published reports of experiments by Jacques-Louis Nauche (1776–1843) and (especially) Nicolas Gautherot (1753–1803) that recalled phenomena he had studied during the summer of 1801 which exhibited contrasting effects of positive and negative galvanic electricity, especially on the body and its sense organs.92 From that work he had come to experience on his own body “[t]he complete reversal of all phenomena, feelings, sensations, etc., at the moment of separation” upon taking himself out of the galvanic circuit.93 That in turn had led him to study the reversal in the action exerted by wires on nonliving systems after they had been removed from the galvanic circuit. For example, the end of the connecting wire attached to the hydrogen-producing end of the pile was seen to elicit the production of oxygen when the wire was tested after its removal from the circuit. Comparable experiments with such wires revealed a comparable reversal of physiological effects. In new experiments from early December 1802 Ritter thought to replace the wires with plates and to enhance their action by using several in series.94 The result was a column composed of plates of a single metal separated by moistened pieces of cardboard; inactive by itself, it was rendered galvanically active by being connected with a conventional galvanic pile.95 Before the end of the month—by which time Ørsted was in Paris—Ritter had drafted the start of a memoir on what he then referred to simply as Säule A, “column A.”96 Although the surviving evidence is not conclusive, it is quite possible that it was in fact Ørsted who coined the term Ladungssäule (Ladningsstøtte and Ladningssøile in Danish); it was certainly he who introduced the terms pile à charger and pile secondaire into French.97 Making what was to be an important connection, Ørsted met Biot on 11 March 1803 at a meeting of the Société Philomatique.98 At the time principally concerned with Ritter’s experiments on light, Ørsted first reported on “Ritter’s experiments with Volta’s column” to the society on May 13, at which time Biot asked him to invite Ritter to apply for the galvanism prize to be awarded by the Paris Academy.99 Ørsted probably knew nothing then about Ritter’s storage

92

93 94 95 96

97

98 99

For a fuller account of this work, plus references, see Ritter 1803a, pp. 101–105, 107 = 1806a, 3, pp. 101–104, 106–107. For good historical accounts of Gautherot’s and Ritters work see Rosenberger 1882–1890, 3 (1889–1890), pp. 117–118, and (especially) Ostwald 1896, 173–180. See also Meyer 1920, xxx–xxxi, Schimank 1933, 190–192, and Ritter 1968, 90–94 (with commentary by Armin Hermann). Ritter 1803a, p. 105 = 1806a, 3, p. 104. Ritter 1803a, p. 114 = 1806a, 3, pp. 112–113. For a modern explanation of Ritter’s experiments see Euler 1981, pp. 28–31. For the dating and the introduction of his terminology see Ritter 1803a, pp. 101, 115 = 1806a, 3, pp. 101, 113. He generalized this terminology by speaking further of “eine Säule A” and “Säulen A” (120, 122, 188; 121–126, 184, 188 bzw. 118, 120, 133; 118–124, 129, 133). For the Danish terms see Ørsted’s letter to Manthey of 4 September 1803, in Ørsted 1870, 1, 156, 157. The larger issue will be addressed presently. Ørsted, travel diary, entry for 11 March 1803, in Ørsted 1870, 1, pp. 121–122. Ørsted, travel diary, entry for 13 May 1803, in Ørsted 1870, 1, pp. 136–137 (quote on 136).

ØRSTED’S PRESENTATION OF OTHERS’

295

column work: its first mention in the surviving letters, in language suggesting Ritter was informing his friend of it for the first time, came in a long letter of 20–23 May 1803 touching upon many topics, in which Ritter noted that it had been a long time since either had written the other.100 Ørsted received the letter on 1 June and replied the next day.101 Ritter then scrambled to complete a paper for submission for the prize, the German manuscript of which he appears to have sent Ørsted on 20 June.102 He also submitted essentially the same paper to Johann Heinrich Voigt (1751–1823), professor of mathematics and physics at Jena and editor of the Magazin für den neuesten Zustand der Naturkunde, where it was published as the third (unnamed) section of a five-part paper (dated July 1803) under the generic title “Experiments and Remarks Concerning Galvanism” in the August and September issues of the journal.103 Thus as much as a month might have separated Ørsted’s receipt of the manuscript and Ritter’s submission of the text to his editor. On 16 August Ørsted presented his substantially reworked French translation of Ritter’s paper to the Institut National, after which he saw to its publication first in Nauche’s Journal du Galvanisme and then in Delamétherie’s Journal de Physique, both under the more revealing title “Experiments on a Device That Can Be Charged with Electricity by Means of Volta’s Electrical Column.”104 Around the time of that meeting Ørsted got word (in letters that have not survived) of Ritter’s discovery of the earth’s electrical poles, the report of which he attached as a “Postscript” and separate “Addition” to the original memoir. From 21 to 31 August Ørsted exhibited Ritter’s experiments before Biot and the galvanic commission, thereafter on 4–9 and 11–13 September collaborating with Biot and Coulomb on a further series of experiments testing Ritter’s claims about the earth’s electrical poles, which Ørsted regarded as Ritter’s most important discovery.105 Alas the commission, whose report Biot read at the Institute on 17 October, could not confirm the claim, though it did substantiate the validity of Ritter’s pile secondaire.106 Although he did not win the prize, that Ritter received a hearing in Paris at all was due solely to Ørsted’s indefatigable intercession on his behalf.

100 101

102

103

104

105

106

Ritter, letter to Ørsted of 20–23 May 1803, in Ørsted 1920b, 2, p. 32. Ørsted, travel diary, entry for 1 June 1803, in Ørsted 1870, 1, p. 142; Ritter, letter to Ørsted of 15 June 1803, in Ørsted 1920b, 2, p. 41. See Ritter’s letters to Ørsted of 15 and 20 June 1803, in Ørsted 1920b, 2, p. 41, 44. From a letter to his publisher of 29 July 1803 it appears that Ritter was still feverishly employed with this work, and expected to remain so until around 6 August (Ritter 1988, p. 131). In a footnote to the reprinting of this article (i.e. Ritter 1803a) Ritter reported that the contents of the third section were an extract (Auszug) of the paper he submitted to the Institut National in the summer of 1803 in competition for the galvanism prize (Ritter 1806a, 3, p. 95). PV, 2 (1912), 692 (meeting of 28 Thermidor, an XI = 16 August 1803). Ørsted reported giving the paper to Delamétherie (letter to Manthey of 25 November 1803, in Ørsted 1870, 1, p. 164). Since Ørsted 1803f contains an “Addition” not present in Ørsted 1803e, the former could not simply have been taken over from the earlier publication. Ørsted, travel diary, summary entries for August and September 1803, in Ørsted 1870, 1, pp. 151, 154–158; Ørsted, letter to Manthey of 4 September 1803, ibid. pp. 155, 157 (the latter for Ørsted’s judgement). Biot 1803, pp. 12–15, 17. Their conclusion was clear, if carefully qualified: “Thus, insofar as one can rely on negative experiments, it would appear that the terrestrial globe does not have electrical poles” (15).

296

K. L. CANEVA

After leaving Paris around the end of October, Ørsted visited Jean-Baptiste van Mons (1765–1842) in Brussels around the second or third week in November.107 It was solely on the basis of what he had learned from Ørsted that Van Mons published an announcement of Ritter’s work in the journal he edited, the Journal de Chimie et de Physique.108 Passing through Haarlem, Ørsted exhibited Ritter’s experiments to the initially unbelieving Martinus van Marum (1750–1837) on 6 and 7 December.109 Van Marum published letters concerning their findings in Van Mons’ and Delamétherie’s journals.110 Thus by around the beginning of 1804 all information published in French concerning Ritter’s work on the storage column had come directly or indirectly from Ørsted. Indeed, Ritter never published in German a report of his discovery of the earth’s electrical poles such as appeared as additions to Ørsted’s French memoir.111 Although Ritter expressed his intention of writing directly to Van Mons, Delamétherie, and Gilbert (editor of the Annalen der Physik), in the event he largely left it to Ørsted to publicize this work.112 We thus find Ørsted in familiar roles: the personal representative and handson promoter of another’s work in a major scientific center (Paris and Berlin); the primary presenter of another’s work in a more accessible published form and language; the traveler who spread the word through his personal contacts in several countries; and the coiner of new terms. Without Ørsted’s intercessions, it is possible that Ritter’s device would not have received its handy designation, and it is all but certain that knowledge of it would have remained much more restricted. With the foregoing framework, we can now more easily consider in closer detail the substance of Ørsted’s presentation of Ritter’s work. Here, as elsewhere, Ørsted subtly transformed what he was ostensibly representing even as he remained essentially faithful to Ritter’s ideas. Of Ørsted’s French translation of his “Preisabhandlung”

107

108

109

110

111

112

Ørsted, letters to Manthey of 6 October and 25 November 1803, in Ørsted 1870, 1, pp. 160, 163– 167. Mons 1803, 200. I follow the Dutch practice of lower-casing “van” in the full name, capitalizing it when it stands alone before the last name, and omitting it for purposes of alphabetization. Ørsted, letter to Manthey of December 1803, in Ørsted 1870, 1, pp. 174–176. On 3 December 1803 Ørsted spoke briefly with Jan Hendrik van Swinten (1746–1825) in Amsterdam (Ørsted, letter to Manthey of December 1803, in Ørsted 1870, 1, p. 174). In a letter to Ørsted of 26 December 1803, Ritter thanked his friend heartily for the connections he’d made for him “with Delamethérie [sic], Van Mons, Van Marum, and especially with Van Swinden,” whom he particularly esteemed (Ørsted 1920b, 2, p. 46). Marum 1803a, 1803b. Van Mons had solicited Van Marum’s “summary of the experiments that M. Ørsted [sic] had proposed to do on his passage through this city” (Marum 1803a, 212). Van Marum himself was primarily interested in the “new proof of the identity of the fluid set in motion by the pile with that excited by the ordinary apparatus” afforded by the demonstration that machine electricity produced the same effect on wires as did the voltaic pile (Marum 1803a, p. 213; cf. 1803b, pp. 471–473, 1804, pp. 212–214). In a footnote to the reprinting of his 1803 paper Ritter reported that the “Postscript” and the “Addition” appended to Ørsted’s translation contained “various experiments of mine which I have up to now not yet published [bekannt gemacht] in Germany and which deal especially with this new (electrical?) polarity of the earth that I discovered in August 1803” (Ritter 1806a, 3, p. 96). Ritter, letter to Ørsted of 15 February 1804, in Ørsted 1920b, 2, 57; Gilbert, footnote to Brugnatelli 1806b, 204–205. For another expression of Ritter’s unfulfilled intentions see his letter to Ørsted of 20–23 May 1803, in Ørsted 1920b, 2, p. 32.

ØRSTED’S PRESENTATION OF OTHERS’

297

Ritter wrote to him that “you’ve frenchified it very well [recht gut französirt], and I’ve learned a lot from it.”113 Ørsted himself reported that “Ritter afterwards declared that he understood [it] better than his own.”114 But before considering some of the notable differences between the two versions, we return to the matter of terminology: Who said what when? Ritter reported that half of the memoir in question—the otherwise untitled part III of the wider-ranging “Experiments and Remarks Concerning Galvanism”—had been written in December 1802, while the paper as published bore the overall date of July 1803.115 It is possible that the parts published in the August and September issues of Voigt’s journal—consisting of 25 and 20 pages, respectively—represent those earlier and later sections, although the paper is reasonably continuous across the break. There does, however, appear to be a break in the flow of the paper at §23, five pages before the end of the paper. As mentioned earlier, in this paper Ritter consistently referred to his new device simply as Säule A, and the detailed table of contents to the paper (printed at the end of the August issue) referred initially simply to experiments “With columns in which a number of plates alternate with moist conductors,” without employing any particular noun.116 Only with the entry corresponding to §23, referring to the “Enlargement of the above storage column [Ladungssäule] into an intensification device [Verstärkungsapparat] for Volta’s battery,” does the new term appear.117 By itself that appearance is not particularly surprising, since throughout the paper Ritter had consistently spoken of his device as being charged (geladen) and discharged (entladen), as having a greater or lesser charge (Ladung) resulting from the process of charging (Proceß des Ladens).118 The term was thus a natural one, albeit one he was surprisingly 113

114 115 116

117

118

Ritter, letter to Ørsted of 15 February 1804, in Ørsted 1920b, 2, p. 55. It would appear that Ritter had only recently seen the translation in the Journal de Physique. Ørsted 1828, p. 525. Ritter 1803a, p. 101 = 1806a, p. 101. Ritter 1803b, p. 176 = 1806a, 3, p. 97. Of the six papers listed in this issue’s table of contents, only Ritter’s has a detailed, four-page analysis of its contents, the other listings being limited to title and author. Such detail thus does not appear to have been the editor’s practice. It is also clear that the table of contents was not prepared directly from the paper as published: sections 4 and 5 as described in the table of contents apply to section V of the paper as printed in the issue for the following month, whereas section IV of the published paper is not captured by the items listed in the table of contents (and thus were likely subsequently inserted). The same discrepancy was uncorrected in the reprinting of the table of contents with the reprinting of the paper in Ritter 1806a. Since the table of contents was printed at the end of the issue, it is possible that it was composed and/or printed after the memoir itself had gone to press. Ritter 1803b, p. 177 = 1806a, 3, p. 97. Ritter’s reference at one point to “meine Säule” (200 bzw. 143) made me wonder whether his terminological reticence might betray a desire to leave that space open for others to fill with something along the lines of “Ritter’s pile.” Indeed, Bernoulli appended to his use of the term “pile à charger” the suggestive clause, “which, in commemoration of its author, one should call colonne rittérienne (Bernoulli 1804, p. 135 [italics in original]∼ 1805, p. 100 [“the Ritterian pile”]). Gilbert omitted this suggestion from the German extract he published (Bernoulli 1806). The usage is so pervasive as not to require citation of particular pages. Schimank (1933, pp. 191– 192) explained Ritter’s likening the operation of his storage column to the charging of a Leyden jar with reference to his earlier work on dry piles, quoting without citation passages from Ritter 1802a (815, 817, as cited from 1806a, 2, 273–274, 276) that can be found in Ritter 1968 (85, 87). Without employing any particular name for his device, Ritter first informed Ørsted of his work in the following terms (letter of 20 May 1803, in Ørsted 1920b, 2, p. 34):

298

K. L. CANEVA

slow to adopt. It is, however, curious that he began what appears to be a late addition to the published paper with an impersonal construction that suggests he was (now) responding to someone else’s perception of the nature of his device: 23. One has seen in my column A an electrical charging device [electrischer Ladungsapparat], which, with regard to the quantity of electricity it receives in a given time, exceeds the largest known electrical battery just as much as Volta’s galvanic [battery exceeds] the largest known electrical machines with regard to the quantity of electricity it furnishes in a given time. There remains meanwhile the question whether that column—since it is a charging device, and one of uncommon capacity—cannot be taken to the point where it stands in precisely the same relationship to the voltaic battery, and performs the same functions, as the Leyden jar and its extension (the electrical battery) [stand] to the electrical machine, so that it would thus at the same time become a true intensification device for Volta’s battery.119

Postponing for the moment pursuit of the substantive lead thus opened up, I would here underscore the significance of Ritter’s intimation that his perception of the nature of his new pile was influenced by someone else’s. From what little is known about the contacts Ritter was then maintaining, Ørsted would appear to be the likeliest person to have played that role.120 The matter must remain undecidable, however, given the complete loss of Ørsted’s letters to Ritter and what appears to be the loss of one or more of Ritter’s letters to Ørsted (on which more below). Nevertheless, what one can say with certainty is that in a letter to Gilbert of 22 August 1803 Ritter referred only in general descriptive terms to his “construction of apparatus that are for the Voltaic pile the same thing that the Leyden jar was in its day for the electrical machine,” without giving his new device a name.121 On the

I make ungalvanized [ungalv.] columns out of 50, 100, 200, etc., times copper and water, without a third body. They yield nothing. I connect them with galvanic columns for 3–5 minutes, and now they produce for a period of time chemical action and shocks and sparks, etc.; eventually it subsides. These columns exceed the charging capacity [Ladungscapacität] of Teyler’s battery many thousands of times. I make columns of this kind that produce purely chemical action and no physiological, others that produce purely physiological and no chemical. With the same quantity of moist conductors it depends solely on the number of interruptions by copper, etc.

119 120

121

In a long letter of 28 March 1803 to his sometime patron, Ernst II, Herzog von Sachsen-Gotha und Altenburg (1745–1804), Ritter emphasized two aspects of his work: (1) He announced “the communication of the galvanic action from an ordinary voltaic column to a device” whose construction and operation he described in some detail, though without the language of “charging” (Poppe 1972, pp. 188–189). Without adopting any particular name for his device—which would be for galvanism what the Leyden jar is for electricity (189)—he once referred in passing to “the connection of the copper column [Kupfersäule] with the voltaic” (190). (2) He announced a fundamental distinction between the action of voltaic columns on organic and inorganic bodies depending on the size and number of plates (189–190). Ritter 1803a, pp. 196–197 = 1806a, 3, p. 140. Another possibility might be Duke Ernst II, at whose home Ritter had performed galvanic experiments during January and February 1802. It is highly unlikely that Ørsted would have been aware of Ritter’s as yet apparently unique use of the term Ladungssäule in the as yet unpublished table of contents. In a long letter to Manthey of 4 September 1803 reporting on Ritter’s work and his own efforts on Ritter’s behalf, Ørsted followed his description of the composition and operation of Ritter’s new pile with an impersonal construction notably noncommittal with regard to the origin of the new term: “It can therefore also be called Ladningsstøtten” (Ørsted 1870, 1, p. 156). Ritter 1803c, p. 106.

ØRSTED’S PRESENTATION OF OTHERS’

299

other hand, at least as of 16 August 1803, when Ørsted presented his “frenchified” version of Ritter’s paper to the Institute, he was speaking regularly of Ritter’s device (“la pile de Ritter”) as either a “pile à charger” or a “pile secondaire,” the latter coinage paired with the designation of the original voltaic pile as the “pile primitive.”122 Throughout that memoir and its two appendices Ørsted used pile à charger and pile secondaire frequently and essentially interchangeably. Biot, in his commission report of 17 October, identified the device under investigation as one “que M. Ritter nomme pile secondaire,” invoking a term Ritter himself never used but which Biot continued to employ as the unproblematic standard.123 That Ørsted continued to use both terms—at least for a while—is demonstrated by the fact that for Van Mons what Ritter had discovered was a pile à charger, whereas Van Marum, like Biot, identified the device inaccurately as one “que Ritter appelle pile secondaire,” both men having derived their information directly from Ørsted.124 Both men’s reports appeared in the same issue of Van Mons’ Journal de Chimie et de Physique (No. 14, nominally dated 15 Brumaire, an XII = 7 November 1803, though clearly published later). Two issues later (No. 16, nominally dated 15 Nivôse, an XII = 9 January 1804) there appeared a note by Luigi Brugnatelli (1761–1818), professor of chemistry at Pavia,. in which he conveyed Volta’s criticism of Ritter’s interpretation of the operation of his designedly so-called “pile à charger” in terms of a battery-like charging in favor of chemical changes produced by the action of the primary voltaic pile.125 Ørsted defended Ritter’s understanding of the operation of his Ladungssäule in a published letter to Gehlen of 4 March 1806, prompting Ritter to acknowledge his friend’s intercession on behalf of what he, too, referred to simply as the Ladungssäule.126 Having been remarkably slow to bother with a distinctive term for his new device, Ritter had come regularly to employ the German equivalent of what might possibly have been the originally

122

123

124 125

126

Ørsted 1803e, 101, 103, 101, 101 = 1803f, 347, 348, 347, 347 = 1920a, 1, 216, 217, 216, 216, respectively, to cite only the first appearance of each. Biot 1803, 13. Biot’s influential Elementary Compendium of Experimental Physics used exclusively the language of “pile secondaire” alongside “pile primitive” and “pile ordinaire” (Biot 1817, 1, pp. 565–569 = 1821, 1, pp. 676–681, in the chapter “Des Piles secondaires”). Swiss physicist ArthurAuguste de La Rive (1801–1873) likewise employed “pile secondaire” and likewise attributed the term’s coining to Ritter (La Rive 1854–1858, 2 [1856], 655). Mons 1803, p. 200; Marum 1803a, p. 212. Brugnatelli 1804, p. 132; 1805, p. 490–491 (“Ladungssäule” on 491). Possibly following Van Mons’ explicit assignation—as opposed to Ørsted’s de facto usage–Brugnatelli also explicitly attributed the new name to Ritter. In a letter of 2 January 1804 to Angelo Bellami from Como after he’d spent many hours with Ritter, Alessandro Volta (1745–1827) reported that Ritter “agrees with almost all my ideas, even regarding his secondary piles [sue pile secondarie]” (Volta 1918–1929, 4 [1927], p. 271; for dating cf. Volta 1949–1955, 4 [1953], p. 324). In an undated manuscript note composed on or after 28 May 1804, Volta spoke of the “pile … secondaire que l’auteur [Ritter] a appellé[e] improprement pile à charger, et qu’on pourroit à plus bon droit nommer pile à changer,” though he himself continued to employ the language of “pile secondaire” and “pile primaire” in the remainder of the note (Volta 1918–1929, 2 [1918], 189). Ørsted 1806a, p. 500 = 1920a, 1, pp. 273–274; Ritter, letter to Ørsted of 10 July 1806, in Ørsted 1920b, 2, p. 177. Ørsted himself, in his rendering of Ritter’s Ladungssäule paper, had indicated his acceptance of a chemical explanation of its operation (Ørsted 1803e, pp. 101–102 = 1803f, 347 = 1920a, 1, pp. 216–217).

300

K. L. CANEVA

French—and Ørstedian—pile à charger.127 It is not hard to imagine Ørsted’s feeling the lack of a simple and distinctive name as he attempted to render Ritter’s ideas into a French text that clearly foregrounded the device. And if he already had had available to him a term sanctioned by Ritter himself, it seems less likely he would have felt the need for the alternative coinage “pile secondaire.” In harmony with what had since become general usage in French, Ørsted used only his term, “pile secondaire,” in an unpublished autobiographical sketch prepared in or after 1847 for a French biographical dictionary.128 The fortunate fact (cited above) that Ritter himself recorded that Ørsted’s French version of his paper was based on the German text published in Voigt’s Magazin makes it possible to judge the extent to which Ørsted transformed Ritter’s work, both by way of clarification (as Ritter recognized) and by way of selective emphasis. Here, at least, well-grounded analysis can replace speculative reconstruction. As usual, Ørsted was in the main quite faithful to his source: there can be no doubt that his intention was to represent Ritter’s work as clearly and accurately as possible. Certainly that is true with regard to what one would think of as the heart of the paper, the construction and operation of Ritter’s pile à charger, “a device that can be charged with electricity by means of Volta’s electrical column,” as the French title made clear. But that focusing—not just in the title, though

127

128

Cf. Ritter, letters to Ørsted of 15 February 1804 and February 6, 1805, in Ørsted 1920b, 2, pp. 55, 95 (“meine Ladungssäulen”). In response to a request by Van Mons for information on Ritter’s recent experiments, Christoph Bernoulli (1782–1863—who taught at the Pädagogium in Halle from 1802 to 1804 before returning to Basel in 1806–reported on his visit with him in a very interesting letter written (it would appear) during the second half of July 1804. Among other things, Bernoulli witnessed Ritter’s “experiments with the new battery of his invention, composed of a single metal, and which he calls pile à charger” (Bernoulli 1804, p. 134 [italics in original] 1805, p. 99 [“the charging pile”]; this information was omitted from Gilbert’s edited translation of the paper, though the term Ladungssäule appeared elsewhere [Bernoulli 1806, p. 102]). Although one cannot deter mine whether Bernoulli’s letter was composed in German or French, his information at least came from the horse’s mouth. The letter contains several clues that allow it to be dated: (1) Bernoulli reported that Ritter was living in a town near Jena; in a letter to Ørsted of 4 August 1804 Ritter said he had recently moved to Dornburg—a town 10 km northeast of Jena—where he wife was living (Ørsted 1920b, 2, 69). (2) Ritter had just been appointed to the Munich Academy; in the same letter to Ørsted of 4 August 1804 Ritter reported the receipt of a letter fourteen days earlier offering him an academic appointment in Munich (ibid. 70). (3) Ritter was reported to have been working for six months on “a systematic work on galvanism” (Bernoulli 1804, p. 140 = 1805, p. 102); Ritter announced that plan to his publisher in a letter of 30 December 1803 (Ritter 1988, pp. 134–136; cited in Rehm 1971, p. 89), and first mentioned it to Ørsted in a letter of 15 February 1804 (Ørsted 1920b, 2, p. 54). Taken together, these clues point to the second half of July 1804 (notwithstanding the nominal date of the issue, 15 Pluviôse, an XII = 5 February 1804!). Despite Ritter’s general adoption of the language of Ladungssäule, he curiously omitted any mention of it in a discussion of the chemical activity of wires previously connected with a voltaic pile in which he explicitly cited Ørsted’s presentation of his work before the Institut National and in the Journal de Physique (Ritter 1804, pp. 696–697). In editorial comments added to his translation of a paper by Brugnatelli on the “hydrogenation of gold” (Brugnatelli 1806a, p. 310, 1806b, p. 202), where Brugnatelli further argued the case for the chemical nature of Ritter’s pile, Gilbert quoted this passage (somewhat modified, and in the context of a nearly page-long citation from Ritter 1804) in a long footnote to Brugnatelli’s discussion of Ritter’s galvanic polarization of gold coins that Gilbert took to be an implicit reference to Ritter’s Ladungssäule (Gilbert, long footnote in Brugnatelli 1806b, pp. 203–205, on pp. 203–204; cf. Brugnatelli 1806a, pp. 310–311). I have no explanation for Ritter’s curious failure to identify his column by name. Ørsted 1847, [5].

ØRSTED’S PRESENTATION OF OTHERS’

301

importantly there—on the device was much more pronounced in Ørsted’s version than in Ritter’s original presentation. Ørsted entirely eliminated all of the merely personal and historical contextualizing of Ritter’s work in its relationship to that of Nauche and Gautherot that had marked Ritter’s own account. Such decontextualizing is always an important step in the objective framing of a universally valid scientific phenomenon, and it generally requires treatment by a succession of others to divest it of those epistemologically superfluous particulars.129 Ørsted began his account by situating the new device within the rationalized context of electrical science, a context within which Ritter himself only eventually came to situate it: The discovery of the electrical pile offers us a device [appareil] that, by the quantity of electricity it disengages, surpasses the strongest machines heretofore imagined. This discovery invited another: it was necessary to search for a device capable of receiving as much electricity as that of Volta could produce. That is what the celebrated Ritter, to whom physics is indebted for so much brilliant work [tant de lumières], has succeeded in doing. He has found a chargeable device [appareil à charger] whose electrical capacity surpasses that of the largest electrical battery as much as the productivity of the pile surpasses the largest electrical machine.130

But before continuing with a description of the new device, he recalled a prior observation of Ritter’s that formed the basis of the present work, namely that “an animal body that has been for a time in the galvanic arc passes, on leaving it, from the state it was in when it formed part of the arc into the opposite state;. … He found, more than two years ago, that an inorganic body [la nature inorganisée] is subject to the same law.”131 Both versions, German and French, devoted several pages to a discussion of chemical, physical, and physiological phenomena that Ritter characterized as exhibiting an “Umkehrung” in their effects vis-à-vis the voltaic pile, and that Ørsted characterized as “phénomènes d’inversion,” thereby giving them a more explicit identity as a class.132 Only Ørsted, however, spoke explicitly in terms of a general “law.” (Recall his like-spirited linguistic transformation of Winterl’s notions into “Hauptsätze”). Indeed, he succeeded in passing on that language to Van Mons, who concluded from his recounting of such phenomena that “[o]ne can thus establish as an invariable law [loi constante] that all bodies that have had one polarity in the galvanic circle leave it with the inverse polarity. One will easily explain, on the basis of this law, the phenomena presented by Ritter’s pile à charger.”133 Both versions gave prominent and extended discussion of Ritter’s finding “that the maximum of charge for the chemical action [or effect (Wirkung)] is

129 130 131 132 133

Cf. Caneva 2001, pp. 20–21 and the quotation from Ludwik Fleck on vi. Ørsted 1803e, p. 97 = 1803f, p. 345 = 1920a, 1, pp. 214–215. Ørsted 1803e, pp. 97–98 (quote on 98) = 1803f, p. 345–346 = 1920a, 1, p. 215. Ritter 1803a, p. 105 = 1806a, 3, pp. 104–105; Ørsted 1803e, p. 98 = 1803f, p. 346 = 1920a, 1, p. 215. Mons 1803, p. 202; in the original the second sentence begins a new paragraph. In his letter of 4 September 1803 to Manthey concerning Ritter’s work, Ørsted wrote that “[t]he basis of his discovery is old; it is namely the law he already discovered a few years ago that a body that has been in a certain galvanic state while it was in connection with a galvanic chain or a galvanic battery, passes over into the opposite [state] when it comes out of that connection” (Ørsted 1870, 1, pp. 155–156).

302

K. L. CANEVA

entirely different from that for the physiological (shock, etc.)”—that is, that differently constructed Säulen A (or piles à charger) produce either predominantly chemical or predominantly physiological effects.134 But whereas Ritter assaulted the reader with a confusing mass of particulars relating to an extended series of experiments with ten differently constructed piles, Ørsted contented himself with a more compact and clearer summary of their results, focusing on the directly relevant particulars, which yielded him a somewhat more intelligible conclusion: “The maximum of chemical action is accompanied by a very weak physiological action; the maximum of physiological action is not accompanied by any chemical action.”135 This finding, which seemed to indicate a particular and fundamental difference between living and nonliving bodies, was seen by both as one of the principal fruits of Ritter’s investigations.136 Such reduction and simplification was typical of Ørsted’s reworking of Ritter’s text. Ritter began his exposition by analyzing the behavior of various Säulen A in terms of crudely deployed concepts of Spannung, Capacität, and Ladung, whereby charge was taken to be the product of tension and capacity, though Ritter quite neglected to indicate how either magnitude was to be estimated.137 Ørsted simply omitted this discussion, which offered no ready explanatory insights. Later in his paper Ritter returned to consider the larger significance of the process of charging, which he believed would shed important light “on the true mechanism of conduction—indeed, one could almost say, of all electrical conduction and charging [Leitung und Ladung].”138 From an analysis of the charging of a Säule A through connection with an active voltaic pile and the subsequent discharging of the former via either an external circuit or internally (as it gradually lost its “charge”), Ritter believed he could conclude for “bodies of the first class”—i.e. metals, according to Volta’s terminology—that every act of conduction is preceded by an act of charging, “that that conduction is nothing more than a continuous [process of] becoming charged of the conducting body from the outside, connected with a simultaneous and persisting [process of] becoming discharged of the same body towards the inside, which holds the former in equilibrium.”139 Thus a conducting body shows itself charged externally only insofar as it is incapable of discharging itself fast enough internally. But whereas Ritter assigned great significance to this finding–“[o]ne sees a result which, even if it were the only fruit of the whole previous work, would nevertheless have richly compensated it”—Ørsted quietly ignored the entire business.140 This omission is especially noteworthy, since Ørsted himself was certainly favorably disposed to conceptualizing phenomena in terms of ‘conflicting’ processes.

134 135 136

137 138 139 140

Ritter 1803a, p. 126 = 1806a, 3, p. 123. Ørsted 1803e, pp. 109–110 = 1803f, p. 350 (no italics) = 1920a, 1, p. 220. “[T]he isolation of the different functions of the pile by means of the same pile à charger”—for example, the separation of the “faculté commotrice” from the “action chimique”—was the second of Ritter’s important findings (after the construction of the storage column itself) that Ørsted reported to Van Mons (Van Mons, 1803, pp. 200, 204). Ritter 1803a, pp. 121–126 = 1806a, 3, p. 118–124. Ritter 1803a, p. 190 = 1806a, 3, p. 134–135. Ritter 1803a, p. 191 = 1806a, 3, p. 136. Ritter 1803a, p. 192 = 1806a, 3, p. 136.

ØRSTED’S PRESENTATION OF OTHERS’

303

One suspects he was induced to tone down the speculative component of Ritter’s paper in deference to what he knew was a strong antispeculative current among his French colleagues. Another notable example of Ørsted’s simplification and clarification of Ritter’s material was his recasting of Ritter’s very difficult-to-follow account of the behavior of ten differently constructed piles, which Ritter cited in support of the rather imprecise generalization “that one and the same quantity of metal and moisture in the circuit [Kreiße] of the battery conducts extremely differently according as metal and moisture are differently distributed among themselves.”141 Ørsted’s presentation of the same experiments is much easier to follow, and he extracted from the results the more specific generalization that “[o]f all the ways of arranging a certain number of both solid and liquid conductors, the arrangement where there are the fewest alternations is the most favorable to the propagation of electricity.”142 A footnote to the next paragraph, in which he elaborated this conclusion, shows Ørsted again disposed to speak in terms of “laws” where Ritter had not ventured to go: The law proven by this experiment accords perfectly with other well-known laws of nature. Light, for example, is much better conducted [conduite] by a transparent and continuous body than by the same body divided into sheets or reduced to powder. I am persuaded that heat is subject to the same law; I shall not delay performing the necessary researches on this subject. It is remarkable that there is still another law of propagation that the electricity of the pile has in common with light, that is, that the action of the pile is also propagated in straight lines; at least it is certain that a metallic wire loses its conductive ability from a bend, and that one can transform a wire of iron, one of the best conductors we have, into a bad conductor by bending it into a zigzag shape. Ritter demonstrated this experiment to me a year ago.143

Although Ritter was as open to drawing analogies as Ørsted, it was in fact Ørsted and not Ritter who made these connections and who dressed up his clarified statement of Ritter’s findings as a law of nature. Ritter devoted several pages to an analysis of the effect of varying the size of the plates in his Säulen A. What he found surprised him. Whereas increasing the plate size of a “voltaic battery” was known to enhance its production of sparks and attendant combustion phenomena and to diminish its chemical and physiological effects, increasing the size of the plates in his column (“meine Säule”) led to an increase in all manner of effects—chemical, physiological, and production of sparks.144 This was not the only point, however, that Ørsted emphasized in his more perspicuous presentation of those experiments.145 To be sure, “those experiments” now went beyond Ritter’s original examination of the behavior of two Säulen A, one with large, the other with small plates, to include a third pile secondaire, also with large plates but with thicker pieces of moistened cardboard between them, which produced the full range of effects to an even more marked

141 142 143 144

145

Ritter 1803a, p. 123–124 = 1806a, 3, p. 121. Ørsted 1803e, p. 106 = 1803f, p. 349 = 1920a, 1, p. 219. Ørsted 1803e, pp. 107–108 = 1803f, p. 350 = 1920a, 1, p. 219. Ritter 1803a, pp. 197, 200 = 1806a, 3, pp. 141, 143; cf. Ørsted 1803e, p. 145–147 = 1803f, p. 358 = 1920a, 1, pp. 227–228. Cf. Ørsted 1803e, pp. 145–149 = 1803f, pp. 358–360 = 1920a, 1, pp. 227–229.

304

K. L. CANEVA

degree than the original pile secondaire with large plates but thinner cardboards.146 The ensemble of these experiments now seemed to hold out the prospect of a virtually unlimited increase in the power of large piles secondaires. Ørsted’s conclusion, absent from Ritter’s original text, payed homage along the way to three French colleagues whose acquaintance he had recently made: The experiments of Fourcroy, Vauquelin, and Thénard [sic] have shown that the chemical efficiency of the electric pile is not increased by a greater width of the plates. Thus the pile à charger is the only device that leads one to hope for an electrical action of great size with regard to both intensity and quantity; for an electric pile of small width will suffice to charge a very large secondary pile.”147

Ørsted repeated (in slightly recast form) Ritter’s “Schema” exhibiting the historical significance and conceptual location of the new device.148 Ritter divided the history of electricity into two periods, the era of insulators and the era of conductors, the former comprising first the discovery of the generation of electricity by insulators (e.g. rubber and amber), then its reception or storage (Aufnahmne oder Ladung) by insulators (e.g. Leyden jar), the latter comprising first the discovery of the generation of electricity by conductors (“Volta’s battery”), then its reception or storage by conductors (“meine Säule”).149 Ørsted grafted his reworking of this material onto an analogous division Biot had made the year before (in the report announcing Napoleon’s galvanism prize) in the history of electricity according as electricity is generated by friction or by the contact of dissimilar conductors.150 He also subdivided the production of electricity into the discovery of the basic phenomenon and the invention of a “machine” capable of producing it (the pile électrique de Volta for the era of conductors), whereby the third moment under each era became the “[d]iscovery of a device appropriate to receive and reinforce the electricity” (first the Leyden jar, then the pile à charger de Ritter).151 What Ørsted hoped to achieve for Ritter’s storage column was conceptual and terminological parity with Volta’s pile. Yet notwithstanding his general enthusiasm for Ritter’s work, Ørsted did not convey to his French readers the unbridled expectations with which Ritter closed his paper: “[O]ur previous views on Volta’s battery [and] its phenomena, on

146

147 148 149 150

151

Ørsted 1803e, pp. 147 = 1803f, pp. 358–359 = 1920a, 1, p. 228. Note that Ørsted’s inclusion of data not present in the original German memoir (as represented by Ritter 1803a) implies that he must have received additional information from Ritter by letter. A comparison of Ritter 1803a, pp. 199–200 (= 1806a, 3, pp. 141–143) with Ørsted 1803e, p. 145 (= 1803f, p. 358 = 1920a, 1, p. 227) indicates the change in focus brought about by the new data, in particular a sharper focus on the potential practical advantages of Ritter’s new device. Ørsted 1803e, pp. 149–150 = 1803f, p. 360 = 1920a, 1, p. 229. On Ritter’s general preoccupation with schemata see Caneva 1997, pp. 46–47. Ritter 1803a, pp. 195–196 = 1806a, 3, pp. 139–140. Ørsted 1803e, p. 154 = 1803f, p. 362 = 1920a, 1, p. 231. Cf. Biot 1802, pp. 3–4 = 1804, p. 235. Ørsted might have seen the separate printing of the report that the commission directed be published (Biot 1802, p. 7 = 1804, p. 238; PV, 2 [1912], p. 518, for the session of 11 Messidor, an X = 30 June 1802). According to Crosland (1967, 21), Biot’s report was published in many journals. Ørsted 1803e, pp. 154–155 = 1803f, p. 362 = 1920a, 1, p. 231. Note the subtle change with regard to the devices’ characterization in terms of ability to receive a charge (Ladung) with Ritter or to reinforce (renforcer) the electricity with Ørsted. That change in emphasis was possibly a response to the implications of the series of experiments that, absent from Ritter’s original paper, he appears to have notified Ørsted of by letter as his friend was at work on his translation.

ØRSTED’S PRESENTATION OF OTHERS’

305

electricity [and] its generation, conduction, [and] charging, [on] its action on both dead and organic bodies, [on] the variously modified irritability of the latter thereby [produced], [on irritability] itself, etc., in general stand before nothing less that a complete reform.”152 Ørsted’s published expectations were more modest.153 New to the French version, as indicated above, was Ørsted’s excited reporting of Ritter’s experiments demonstrating that the earth has electric poles, as revealed initially by the regular orientation of a freely suspended pile secondaire.154 Having received the earliest report in a letter that arrived just as he was preparing to deliver his memoir to the Institute, he was subsequently informed of further experiments—some employing a galvanically charged gold wire—that confirmed the original finding.155 Alas, as already noted, the Institute’s commission, working closely with Ørsted himself, was unable to substantiate these claims. Nevertheless, Ørsted’s faith in his friend’s experiments appears not to have been much if at all shaken, as he reported them to Van Mons with the assurance that Ritter had repeatedly confirmed them.156 Although Christensen has claimed that “[t]he absence of evidence of the galvanic poles experiment overshadowed the successful demonstration of the charging pile, and ruined the entire expedition”—i.e. Ørsted’s trip to Paris, which Christensen interpreted as essentially an effort to vindicate Ritter’s and Ørsted’s brand of dynamical philosophy in Paris–I am aware of no evidence that Ørsted’s Parisian colleagues, so many of whom he had interacted with so often and at such close quarters, came away from the episode with a seriously negative impression of their youthful Danish colleague.157 After all, the storage column did work as announced, Ritter had succeeded in identifying chemically active rays beyond the violet end of the solar spectrum, and Ørsted was only passing on someone else’s claim about the earth’s electric poles—a claim the Institute’s galvanic commission took very seriously, and which they carefully avoided declaring to be false, simply not proven. Winterl’s work was, to be sure, quickly shown to be untenable through the work of Guyton de Morveau in particular—and he got his Winterl directly from the Prolusiones, not from Ørsted’s Materialien—but Ørsted’s efforts on Winterl’s 152 153

154 155

156

157

Ritter 1803a, pp. 200–201 = 1806a, 3, p. 144. Evidence of Ørsted’s desire not to appear too speculative before a French audience is his private remark to Manthey, apropos of Ritter’s notion of the separability of the chemical from the physiological effects of the pile, that “[t]he theory that ascribes to a fine matter all the electrical actions [Virkninger] appears therefore to longer able to survive, since one can thus separate the one phenomenon from the other” (letter of 4 September 1803, in Ørsted 1870, 1, p. 156). Ørsted 1803e, pp. 156–157 = 1803f, p. 363 = 1920a, 1, p. 232; from the “Post-Scriptum.” Ørsted 1803f, p. 364 = 1920a, 1, p. 233, from the “Addition” absent from Ørsted1803e. In this addition Ørsted used the term pile secondaire three times as often as pile à charger. Mons 1803, p. 206. Bernoulli (1804, 141 = 1805, p. 103), in his letter to Van Mons from around July 1804, reported Ritter’s experiments with gold wire, though apparently without having witnessed them himself. Gilbert (in Bernoulli 1806, p. 103) quoted this passage in a very free German translation in order to demonstrate the kind of false impression foreigners were being given about the bizarre beliefs of German physicists. Christensen 1995, p. 165. I know of no evidence—from his letters, diary, or publications—that Ørsted’s purpose in going to Paris in 1802 was “to convince the National Institute that the era of mechanical philosophy was running out” (164), nor was it the case that Ørsted “shar[ed] the zeal of Ritter and the Romantic circle of poets and philosophers rather than the company of fellow scientists” (164). As his diary makes abundantly clear, he spent the lion’s share of his time precisely with fellow scientists.

306

K. L. CANEVA

behalf were mostly directed to a German audience, and it is not clear that his Parisian reputation suffered overmuch over this matter.158

5. ØRSTED’S PRESENTATION OF SEEBECK’S EXPERIMENTS ON THE MAGNETIC POLARIZATION OF METALS BY TEMPERATURE DIFFERENCE Perhaps the most significant, and long-lasting, example of Ørsted’s role as presenter of others’ work was his dissemination of Thomas Johann Seebeck’s (1770– 1831) discovery of what has come to be known as thermoelectricity, the term coined by Ørsted. The identification—both conceptual and terminological—of just what Seebeck had discovered was a complex process in which Seebeck himself played only a late and ineffectual role. Because of his nearly four-year delay in publishing his findings, conceptual and terminological possession of the new field fell to others as word of his experimental finding leaked out and then attracted the attention of a spate of investigators throughout Europe. It was Ørsted who played the pivotal role both in disseminating prepublication information about Seebeck’s work and in shaping its interpretation.159 Once again it was an extended trip abroad—this time from early November 1822 till early October 1823—that made it possible for Ørsted to play that role.160 Having been in Berlin since the end of November, by 2 December he had already spent two mornings and an afternoon with Seebeck being shown the experiments that Ørsted immediately regarded as continuations of his own discovery of electromagnetism.161 Before having the opportunity to undertake his own experiments, Ørsted passed through Jena, where he informed Johann Wolfgang Döbereiner (1780–1849) of Seebeck’s work.162 After repeating them for himself, Döbereiner in turn informed Gilbert, editor of what was essentially the house organ of German physical scientists. In a hodgepodge letter of 12 January 1823 dealing with a variety of disparate subjects—including a “remarkable change in wood due to lightening” alongside a “repeating of Seebeck’s experiments on magnetic electromotors by means of heating”—Döbereiner wrote: Three weeks ago Prof. Ørstedt [sic] from Copenhagen visited me. Among other things he recounted to me that Dr. Seebeck in Berlin had constructed dry ring-shaped elec-

158

159 160 161

162

To be sure, Irish chemist Richard Chenevix closed his French-language review of Ørsted’s Materialien with the damning remark that “[f]or the glory of the eighteenth [sic] century, it is to be hoped that it will hasten to reject the offering of M. Ørsted and the chemistry of Winterl” (Chenevix 1803–1804, 6 [1804], p. 16 1804, p. 199, 1805, 454 [where the century is changed to the nineteenth]). This section is based on Caneva 2001, pp. 5–10. See the letters published in Ørsted 1870, 2, pp. 29–77. Ørsted, letters of 2 December 1822 and April 4, 1823 to his wife, in Ørsted 1870, 2, pp. 31–32, and 59; and later in print: “The most remarkable of all the discoveries, to which that of Ørsted has given occasion, is no doubt the thermo-electricity, discovered in 1822 by Dr. Seebeck” (Ørsted 1830, p. 576 = 1920a, 2, p. 359). Interpolating from scattered references, one can place Ørsted in Jena around the middle of December 1822, in any event before the 17th of the month (Ørsted 1870, 2, pp. 35, 41, 77).

ØRSTED’S PRESENTATION OF OTHERS’

307

tric circuits [Ketten] consisting of copper and antimony or arsenic, bismuth, and the like, which, when they are heated or only warmed by the flame of an alcohol lamp on the places where the two different metals are soldered to each other, act on the magnetic needle every bit as strongly as a voltaic circuit. I hastened to repeat this interesting observation, and found in fact that when a bismuth rod is soldered together on both ends with a bent strip of copper sheeting … a circuit is formed which obtains a magnetically reactive force merely through the warmth of the hand or fingertips with which one touches one of the soldered places …, so that a magnetic needle placed between the two metals…is brought to an easterly or westerly deviation of 10 to 15 degrees.”163

One cannot be sure whether the title’s identification of the circuits as “magnetic electromotors” was Döbereiner’s or Gilbert’s. Döbereiner himself seems to have regarded the phenomena as essentially magnetic, in that regard perhaps being guided by what Ørsted might have passed on as Seebeck’s view of the thermally induced magnetic polarization of metals. Passing through Erlangen, on 22 December Ørsted spent several hours of an evening with Friedrich Wilhelm Joseph Schelling (1775–1854) and his new wife.164 Twenty-one years later, hosting Ørsted at their house in Berlin, Schelling’s wife recalled her pleasure at Ørsted’s showing her how to construct “a thermoelectric ring in order to demonstrate Seebeck’s then brand-new discovery.”165 Although that encounter has no further implications for our story, it shows yet again the importance of personal connections in the dissemination of unpublished scientific ideas among the broadly conceived scientific community of the day—and it really was a community, albeit scattered, defined and held together in no small measure precisely by these kinds of visits. Of much greater significance was Ørsted’s encounter with Julius Conrad von Yelin (1771–1826), Bavarian Oberfinanzrat and member of the Munich Academy of Sciences, whom Ørsted had met by 28 December 1822.166 Having for some time been interested in exploring the relationships among electricity, magnetism, heat, and light, by the summer of 1821 Yelin had noticed that a red-hot iron rod becomes magnetized in any orientation except perpendicular to the magnetic meridian, and he was investigating Domenico Morichini’s (1773–1836) then almost nine-year-old finding that steel needles can be magnetized by the violet light from the sun. Finding that he could magnetize needles with heat alone, he concluded that in Morichini’s experiments the effect was due not to the violet light per se, but to the heat it generated. But having been unable to magnetize rods of brass, copper, silver, and platinum by means of even very strong heating, he concluded that in all cases the ultimate cause of the magnetization was the earth’s magnetism. He was thus primed to respond to Ørsted’s personal communication of Seebeck’s as yet unpublished findings: “After this preliminary

163 164 165

166

Döbereiner 1823, p. 115. The ellipses replace references to a figure. Ørsted, letter of 28 December 1922 to his wife, in Ørsted 1870, 2, p. 36. Ørsted, letter of July 1843 to his wife, in Ørsted 1870, 2, pp. 179–180 (quote on 180). She misremembered the meeting as having taken place in Nuremberg, whence (it seems) the Schellings, Ørsted, and others repaired for a feast the day after their meeting in Erfurt (Ørsted, letter of 28 December 1822 to his wife and of 10 October 1823 to his sister, in Ørsted 1870, pp. 2, 36, 78). Ørsted, letter of 18 December 1822 to his wife, in Ørsted 1870, 2, pp. 37–38. He was still in Munich on 9 January 1823, the date of a letter to his brother (ibid. p. 38).

308

K. L. CANEVA

work, Seebeck’s new thermelectromagnetic experiments, about which the Philosophical Magazine of 1821 contains the first, albeit very superficial notice and with which I … became more familiar through Professor Ørsted’s presence [here in Munich], had to arouse my attention to a high degree.”167 As soon as Ørsted left Munich, Yelin undertook his own experiments, reporting his first results to the Academy on 11 January 1823.168 In an early write-up of his findings (“New Experiments on the Magnetomotive Property of the Heretofore So-called Unmagnetic Metals”) which bore directly on his prior concerns, Yelin continued to refer to Seebeck’s “thermelectromagnetic experiments” as he went on to claim for himself the discovery “that through unequal heating all bodies acquire magnetomotive properties,” a fact he demonstrated by forming circuits composed of a single metal which, when heated at one place, produce a deflection of an appropriately placed magnetic needle.169 Having detected a magnetic effect—a magnetic polarization—by the appropriate heating of a bar of metal not part of a closed circuit, he noted that “[b]ismuth most strikingly exhibits the polarization occurring in this thermomagnetism of metals,” with a footnote glossing the phrase “Thermo-Magnetismus der Metalle”: “It appears to me that we will thus have to denote in a characteristic fashion this kind of magnetic action, in contradistinction to the already known Ørstedian electromagnetism, on account of its peculiar behavior.”170 What was precisely at stake for him became clearer in several subsequent papers. In Yelin’s terminology, “electromagnetic” referred to the action of the electricity set in motion in the connecting wire of a galvanic circuit on a magnetic needle—whereby he explicitly accepted the explanation in terms of an electric current—while his neologism “thermelectromagnetic” referred to the phenomenon of magnetic deflection produced by heat applied to circuits composed of either two metals (as Seebeck had found) or one (as he discovered).171 Seeking to reduce Seebeck’s experiments to their simplest form, he found that if he held discs of two metals between his fingers, a magnetic needle suspended above them registered a deflection. Wondering whether a single metal would suffice to produce “magnetomotive actions [or effects (Wirkungen)],” he found indeed that bars of all metals become “magnetomotors” as long as they are unequally heated in two places.172 That is, one could produce a magnetomotive effect without (it appeared) a closed circuit carrying an electric current. Furthermore, rods of different cross-sections 167

168 169 170 171 172

Yelin 1823c, pp. 418–419. Yelin’s reference is to Tilloch 1821, a short, one-paragraph notice under the heading “Magnetism” whose only relevant sentences were the first two: “The Prussian State Gazette mentions a highly important discovery, which Dr. Seebeck had communicated to the Academy of Sciences at Berlin, in three different sittings. It was on the magnetic properties inherent in all metals and many earths, (and not in iron only as was supposed,) according to the difference of the degree of heat” (462). Given the vagueness of this information, it is clear that it required Ørsted’s input before Yelin had anything to react to. I have not been able to consult the Allgemeine preußische Staatszeitung, which Yelin also appears to have been ignorant of. Yelin 1823b, 4; 1823d, p. 419. Yelin 1823b, p. 361. Yelin 1823b, p. 363. Yelin 1823c, pp. 418–421. Yelin 1823c, pp. 423–425 (quotes on 425 and 424, respectively).

ØRSTED’S PRESENTATION OF OTHERS’

309

did not all exhibit magnetomotive actions similar to those circularly symmetrical actions displayed by “an Ørstedian connecting wire.”173 It was these differences that prompted him to introduce the term “thermomagnetism,” although he did not thereby “wish to dispute the legitimacy of the effect of the natural electricity aroused by heat and set in motion, or to fail to recognize the phenomenon itself as properly electromagnetic.”174 It is thus clear that Yelin did not intend Thermo-Magnetismus to apply to the full range of phenomena discovered by Seebeck, but only to a new class of actions he had discovered which did not appear to involve a closed electric circuit and in which no trace of free electricity could be detected.175 That was not, however, the way others came to employ the term. For them, thermomagnetism was the preferred designation for what Seebeck had discovered until the 1830s. (The term’s appeal was closely related to the popularity in Germany of theories of transversal magnetism, which sought to explain electromagnetic interaction not in terms of a direct action between an electric current and the magnetism of the compass needle—let alone in terms of Ampère’s electrodynamic theory of magnetism—but in terms of a magnetism produced in the conducting wire that consisted in a number of small magnets running around and in the wire head-to-tail in circles perpendicular to its length.) In the long run, however, it was Ørsted’s rendition that carried the day, one he began publicizing during his long sojourn in Paris. In the French capital since January 1823, Ørsted informed the Académie des Sciences on March 3 of “Seebeck’s new experiments on electromagnetic actions” (as the published title had it).176 Its first sentence continued the subtle process by which Ørsted sought to control the interpretation of those experiments: “Seebeck, a member of the Berlin Academy, has discovered that one can establish an electric circuit in metals without the interposition of any liquid. One establishes the current in this circuit by disturbing the equilibrium of temperature.”177 In noting that “[o]ne can only discover these electric currents by means of the magnetized needle,” he was already ignoring the issue, vital to many others, of the legitimacy of speaking in terms of an electric current when what one observes is magnetic effects.178 173

174

175 176

177 178

Yelin 1823c, pp. 424–426 (quote on 426). In the phenomena of thermomagnetism he also noted a connection between the magnetomotive action of metals, their mode of crystallization, and electricity (Yelin 1823a, p. 10). Yelin 1823c, p. 427; cf. 424, where he concluded “That with this magnetization by heat peculiarities manifest themselves by which it is essentially distinguished from the electromagnetism discovered by Ørsted, for which reason I call it in contradistinction therefrom thermomagnetism.” See also Yelin 1823a, pp. 5–8. Yelin 1823a, p. 11. Ørsted 1823b. Anonymous (1823, p. 315), reporting on the meeting of 3 March 1823, says Ørsted entertained the Academy “with the work that Seebeck has just done on electromagnetic phenomena. (See the preceding cahier.)”—i.e. Ørsted 1823b. The report was quickly translated into German (Ørsted1823c) and English (Ørsted 1823f); Gilbert’s German title spoke rather of Seebeck’s “elektrisch-magnetische Versuche.” The minutes for that date record that Ørsted communicated “the result of divers experiments on the motions of electricity determined in certain metals by differences in temperature” (PV, 7 [1916], p. 455). Ørsted 1823b, p. 199 = 1920a, 2, p. 263. Ørsted 1823b, p. 199 = 1920a, 2, p. 264. Ørsted was aware that Seebeck had another theory about these effects, but it is not clear that he had a very distinct idea of what that theory was (see his letter of 4 April 1823 to his wife, in Ørsted 1870, 2, pp. 59–60).

310

K. L. CANEVA

He went on to suggest several new coinages: “It will from now on doubtless be necessary to distinguish this new class of electric circuits by an appropriate term; and as such I propose the expression thermoelectric circuits or perhaps thermelectric: at the same time one would be able to distinguish the galvanic circuit by the name hydroelectric circuit.”179 He quickly abandoned his short-lived preference, “thermelectric,” in favor of the term that would eventually become the one most commonly applied to this new class of phenomena.180 Having demonstrated the action of the thermoelectric current developed by a single couple, Ørsted thought to see if the equivalent of a voltaic column might be constructed out of serially arranged thermoelectric units.181 He enlisted the collaboration of Joseph Fourier (1768–1830), reporting their joint findings to the Academy on 31 March 1823.182 As far as the French were concerned, what Seebeck had discovered was what Ørsted had represented him as having discovered: thermoelectricity.183 In reporting on Seebeck’s work to Michael Faraday in a contemporaneously published letter of 18 April 1823 André-Marie Ampère (1775–1836) employed without comment Ørsted’s language of electric current and thermoelectric pile.184 In his report of 24 April 1823 on the work of the Academy in 1822, Fourier announced that Seebeck had shown that the contact of dissimilar metals of different temperatures is sufficient “to occasion very noticeable magnetic effects”; he went on to describe his and Ørsted’s experiments “on these thermoelectric actions” (and to speculate on their significance for the explanation of terrestrial magnetism).185 Blainville’s several reports explicitly followed Ørsted’s lead in characterizing the “electromagnetic actions” of the “thermoelectric circuit” as

179 180

181

182

183

184 185

Ørsted 1823b, pp. 199–200 = 1920a, 2, p. 264. After their introduction, Ørsted went on to use thermélectrique six more times, thermo-électrique not again in that paper. His later Danish account already reported impersonally–if not entirely accurately–that “[o]ne now calls the Seebeckian circuit the thermoelectric circuit, and the Galvanic, in opposition thereto, the hydroelectric circuit; in Danish one could call the former the varmeelectriske, the latter the vandelectriske circuit [Kiæde]” (Ørsted 1823e, 9 = 1920a, 2, p. 461). Ørsted believed that Seebeck had not hit upon the idea of trying this arrangement (letter of 5 May 1823 to Prince Christian Frederik, in Ørsted 1870, 2, pp. 70–71). In a preliminary report on 24 March 1823 Ørsted announced that “he has succeeded in augmenting the effects he has designated with the name thermoelectric, that is, which result from the contact of different substances and from the inequality in temperature” (PV, 7 [1916], p. 465). His report of March 31, 1823 on his and Fourier’s experiments in constructing circuits of multiple pairs of metals opened with language reinforcing his view of the subject: “I have had the honor of showing this illustrious assembly the remarkable experiments by which Seebeck has proven that one can establish an electric current in a circuit formed exclusively of solid conductors merely by disturbing the equilibrium of temperature.… We are therefore in possession of a new kind of electric circuits, which one can call thermoelectric circuits in thus distinguishing them from galvanic circuits, which it would be convenient henceforth to call hydroelectric” (Ørsted 1823d, p. 375 = 1920a, 2, p. 272). In this report he used thermo-électrique exclusively, applying it to circuits, currents, and elements (375, 383, 384 bzw. 272, 278). In neither report did he yet use the noun thermoelectricity. In a section entitled “Thermomagnetism becomes thermoelectricity,” Nielsen asserted that it was mainly through Ørsted’s writings that news of Seebeck’s “thermoelectric” work circulated through Europe (K. Nielsen 1989–1991, 21, pp. 391–395). Ampère 1823, p. 398. Fourier 1826, p. 239.

ØRSTED’S PRESENTATION OF OTHERS’

311

“thermoelectric effects” produced by the electricity in the circuit.186 In the article “Thermo-electricity” that Ørsted sent to David Brewster in 1827 for anonymous publication in the Edinburgh Encyclopædia, he brushed past Seebeck’s inclination not “to consider the effect thus produced as a true electrical current, but an effect sui generis,” informing his readers that “Professor Ørsted has proposed to call the current discovered by Dr. Seebeck the thermo-electrical current, and in consequence of this to distinguish the action hitherto called Galvanism, by the name of the Hydro-electrical current. Hence we have now the names thermo-electricty and hydro-electricity, to which we could add the name tribo-electricty for the electricity produced by friction.”187 In his 1828 autobiography Ørsted recorded simply that “[i]n Paris in the company of the mathematician Baron Fourier he carried out new experiments on the thermoelectricity discovered by Seebech [sic].”188 The discovery Seebeck finally announced in the title of his long-awaited memoir of 1825 was the “Magnetic Polarization of Metals and Ores by Means of Temperature Difference.”189 In repeating Ørsted’s electromagnetic experiments with the intention of investigating “the mutual relationship of the electrical, chemical, and magnetic activities in the galvanic circuit,” Seebeck had noticed “that not just the action at the point of contact of the metals with each other, but much more the inequality of the action at both points of contact of the metals with the moist conductor determines the magnetic polarization of the whole closed circuit [Kette],” and he thus suspected “that a magnetic polarization could be produced by any disproportion taking place in the state of the points of contact of two metals connected to each other in a circle.”190 Thus was initiated the series of experiments he performed between the end of July 1821 and the beginning of February 1822.191 Having chosen bismuth and antimony for his first trial because their behavior with copper in a voltaic pile was deviant and variable, he noted a deflection of the magnetic needle when he held samples of the two metals between his fingers. Through further trials he realized that not moisture but heat was the true “cause of the magnetism.”192 He thereupon tried pairwise combinations of 28 different metals and ores joined together to form a ring, heating one junction and noting the direction and rough strength of the resulting “magnetic polarization”

186

187

188 189

190 191 192

Blainville 1823b, pp. 199, 201, 200, respectively; cf. 1823a, lii. Henri-Marie Ducrotay de Blainville (1778–1850) was editor of the Journal de Physique. Ørsted 1830, p. 584 = 1920a, 2, pp. 384–385. Writing in 1823, Ørsted’s friend, Christopher Hansteen (1784–1873), simply took Ørsted’s characterization as Seebeck’s: “Dr. Seebeck called these phenomena thermoelectric, the Ørstedian hydroelectric” (Hansteen 1823, 106). Years later, Auguste de La Rive reported similarly that “Seebeck gives the name thermoelectric to this current and to the couple that generates it in order to distinguish it from the hydroelectric current and couple otherwise called voltaic current and couple” (La Rive 1854–1858, 2 [1856], p. 473). Ørsted 1828, p. 539. As Seebeck explained, the published paper was an “extract” from four lectures delivered at the Academy of Sciences in Berlin on 16 August 1821, 18 and 25 October 1821, and 11 February 1822, plus later additions in the form of footnotes and an addendum (Seebeck 1825a, p. 265 = 1825b, p. 1). This clarification is omitted from Seebeck 1826. Seebeck 1825a, p. 266 = 1825b, p. 2. Seebeck 1825a, p. 265 = 1825b, p. 1. Seebeck 1825a, pp. 270–271 = 1825b, p. 7.

312

K. L. CANEVA

by means of the deflection of a suspended magnetic neede.193 On the basis of further experiments he argued that the magnetic polarization he observed could not be solely due to any free electricity—such as could be detected with an electroscope—produced there, and that one was therefore not justified in calling such circuits “electromagnetic” until such time as one could clearly demonstrate the presence of freely circulating electricity.194 Contrast this reticence with Ørsted’s immediate assimilation of the new phenomena to electromagnetism, which for him (by them) referred to the mutual interaction between magnet and current. Seebeck explained the phenomena in terms of the excitation of a “magnetic polarity” or “magnetic polarization” in the bimetallic circle by means of a difference in temperature.195 Electricity as such played no role. Nor in his original paper did Seebeck employ the language of thermomagnetism Yelin had introduced 14 months after the last of the four presentations on which Seebeck’s memoir was based (that of 11 February 1822). As a noun it occurred only once, in a subsequently added footnote, where he used it matter-of-factly, without any discussion of terminological issues.196 As an adjective, “thermomagnetic” was used four times in the 109-page memoir—all but once in footnotes—applied to the action of metals, to phenomena, and to experiments, including “my thermomagnetic experiments.”197 That was the extent to which Seebeck assimilated his discovery of magnetic polarization to the language of thermomagnetism in this, his first and last published paper on the subject.198 To many of his contemporaries, however—especially among German contemporaries of his generation—that was what Seebeck was credited with having discovered until well into the 1830s.199 Eventually, however, everyone came to adopt both the language of thermoelectricity that Ørsted had introduced and the explanatory conception that underlay it.200 The first beachhead of that usage was established in French publications written while Ørsted was in close personal contact with the leaders of French physics in Paris. 193 194

195

196

197

198

199

200

Seebeck 1825a, pp. 266–283 = 1825b, pp. 2–19. Seebeck 1825a, pp. 296–297 = 1825b, pp. 32–33 (quote on 297 bzw. 33). Right after quoting this passage, Nielsen added that “[f]rom then on, Seebeck talked only about thermomagnetism and lost no opportunity to emphasize that it was not an electric effect” (K. Nielsen 1989–1991, 21, p. 382), although he neglected to cite any specific instances of Seebeck’s use of the term. For a sampling of such usage, see Seebeck 1825a, pp. 312, 330, 334, 338 = 1825b, pp. 48, 66, 70, 74, among many occurrences. Seebeck 1825a, p. 297 = 1825b, p. 33. In the Annalen der Physik version this instance was silently transferred to the body of the paper (1826, p. 142). Seebeck 1825a, pp. 301, 303, 320, 354 = 1825b, pp. 37, 39, 56, 90 (quote on 354 bzw. 90). The first occurrence was incorporated into the body of the paper in Seebeck 1826, p. 145; the second was omitted; the fourth remained in a footnote. The third instance (“thermomagnetische Erscheinungen”) occured in the body of the original paper as printed, but after a dash, suggesting its later addition; it was omitted from 1826, 154, in a recast sentence. In four subsequent unpublished papers, however—read at the Berlin Academy between February 1824 and November 1830—he appears to have wholly adopted the language of thermomagnetism; see Poggendorff (1841, p. xxxviii) for titles and dates. Humboldt 1993, p. 58 (from the first and second lectures, delivered c. 6 and 13 December 1827); Muncke 1836, p. 710; Pfaff 1838, p. 732. Although Poggendorff recognized that, by 1839, the designation “thermoelectric” had become normal, he himself still preferred to identify Seebeck’s discovery as “thermomagnetism” (Poggendorff 1841, pp. xxxiii, xxxvi–xxxvii). A more detailed presentation of this story will be given in Caneva, forthcoming.

ØRSTED’S PRESENTATION OF OTHERS’

313

II. ØRSTED’S PRESENTATION OF HIS OWN WORK 6. ØRSTED’S PRESENTATION OF HIS CHEMICAL THEORY TO GERMAN AND FRENCH AUDIENCES While in Berlin in during the summer of 1812 Ørsted prepared a German exposition of his chemical theory, published in August as View of the Chemical Forces of Nature Derived from Recent Discoveries (hereafter Ansicht).201 In Paris during the winter of 1812–1813—where, he opined, one was like in Turkey as far as German books were concerned—he undertook a substantially recast French translation, published as Researches on the Identity of Chemical and Electrical Forces (hereafter Recherches).202 Ten years later he reflected on some of the differences he had encountered in writing for audiences in the two languages: I have had many opportunities to notice that it is almost impossible to make my [electrochemical] theory comprehensible to the French without also explaining to them some of the features of Naturphilosophie. If in Germany I’m often tempted to declare myself against Naturphilosophie when I see the misuse one makes of it, in France I see myself all the more called upon to go to its defense; or rather I feel a fundamental difference in scientific way of thinking that I wouldn’t have imagined to be so great if I hadn’t so often felt its living presence.… It is good that the nations of Europe have character differences just like people, whereby the one-sidedness is eliminated that would otherwise surely take the upper hand.203

Those words, however, were written at a time when Ørsted had become even more self-conscious about his early association with Naturphilosophie than he had been in 1812 and 1813, and in both German and French versions his indebtedness to that philosophical system—and to the work of Ritter and Winterl that contemporaries would have associated with it—was clearly acknowledged.204 A baseline for comparison of the two versions must be the recognition that all of Ørsted’s fundamental concepts relating to the relationship between electrical

201

202

203 204

Ørsted 1812b; letter of 18 August 1812 to his sister Sophie, in Ørsted 1870, 1, p. 301. It seems likely that Ørsted himself had intended his book to be called Ansichten, as that was the form in which he consistently referred to it (letters of 18 August 1812, 25 February 1813, and August 1843 to his sister, brother, and wife, in Ørsted 1870, 1, pp. 301, 313 and 2, p. 199; 1821a, p. 12 = 1920a, 2, p. 447; 1828, p. 530). Ørsted 1813; letter of 25 February 1813 to his sister, in Ørsted 1870, 1, 315 (“Her i Paris er man som i Tyrkiet, hvad tydske Bøger angaaer.”). Ørsted, letter of 23 February 1823 to his wife (from Paris), in Ørsted 1870, 2, pp. 54–55. Ørsted 1812b, pp. 11–12 (which has “Winter”) = 1920a, 2, p. 41; 1813, 10. In a “Post-Scriptum” to his introduction to the French edition he reiterated his regard for and indebtedness to Ritter and Winterl (1813, 12–13, 17–18 = 1920a, 2, pp. 173, 175). Where in German he wrote “Dankbar erkennen wir die Verdienste der Naturphilosophen unseres Zeitalters, auch mit Rücksicht auf die chemischen Ansichten” (12 bzw. 41–42) the French had “Nous ajouterons enfin que la philosophie de la nature qu’on a cultivée depuis vingt ans en Allemagne, pourrait aussi réclamer ses droits sur quelques vues que nous allons proposer” (11). In his preface, however, Marcel de Serres had maintained “that the philosophy known in Germany under the name of philosophy of nature (natur philosophie), and which holds sway in several parts of this country, will only be able to have a baleful influence on the sciences, especially on those of observation; for whenever one will not proceed from experiment in order to arrive at general ideas one will always risk going astray” (Serres, in Ørsted 1813, pp. xiv–xv).

314

K. L. CANEVA

and chemical forces (as representatives of the behavior of the presumed underlying fundamental forces of nature) and all of his important technical terms in German made it over into French: Urkräfte became forces primitives, the allgemeine Grundkräfte der Körperwelt the forces primitives et universelles des corps, the frequently deployed concept of Wirkungsform as the distinguishing aspect of the various forces became forme d’activité, down to Gebundenheit, which appeared as état de gêne.205 More generally, in both he identified his theory as “dynamical” because it took as its starting point the primitive forces of nature.206 Two of his chemical series arranged substances according to their relationship either to combustion or to acid-base neutralization, properties he termed the verbrennungshervorbringende Thätigkeit (activité ignifère) and the neutralisirende Thätigkeit (activité neutralisante), respectively; the former series further classed substances according to which of the opposing forces predominated, Brennkraft (force de combustibilité) or Zündkraft (force comburente).207 Thus the reader of either book would have come away with knowledge of essentially the same general theory. Nevertheless, some of the more explicitly Kantian notions—in particular, the construction of matter out of opposing fundamental forces, a conception taken over by both Schelling and the early Ørsted—had decidedly less visibility in the French version.208 For example, in presenting his view of the undulatory propagation of the electrical forces, he added in German but not in French “that it is precisely these forces by means of which space becomes corporeal,” just as in summarizing the general significance of the chemical and electrical forces, he added in German but not in French “that the mechanical filling of space can derive from the same [forces].”209 In some cases the language of the German was more redolent of the conceptual baggage of Naturphilosophie than were the corresponding passages in the French. One notable instance was his discussion of what was for him a major unresolved issue in chemistry: how properly to arrange chemical species in appropriate (“more natural”) series.210 The image he conjured up in the Ansicht was of nature as “a whole caught up in ceaseless development [ein in unaufhörlicher Entwicklung begriffenes Ganzes],” where, “eternal in her laws,” she directs the ceaseless transformation of objects and, “out of the forces lying dormant there [aus den darin schlummernden Kräften], gradually develops a creature 205

206 207 208

209

210

For illustrative examples see Ørsted 1812b, pp. 5, 114, 150–151, 82 = 1920a, 2, pp. 38, 88, 104, 73; 1813, 4, 109, 141–142, 79, respectively. Ørsted 1812b, p. 207 = 1920a, 2, p. 129; 1813, p. 197. Ørsted 1812b, pp. 73, 79–80 = 1920a, 2, pp. 69, 72–73; 1813, 73, 78–79. Jacobsen (2000, 149) noted that the “Post-Scriptum” added to the Recherches acknowledged Ørsted’s debt to Ritter and Winterl, “but there is no mentioning of the philosophical influence from Kant and Schelling.” Ørsted 1812b, pp. 140, 150 = 1920a, 2, pp. 99, 104; cf. 1813, pp. 130–131, 141–142. At one place in the later work he did refer to the two fundamental forces of nature as “the basis of all chemical action, as of all mechanical existence” (1813, p. 198). The odd wording would appear to be a poor translation of the German, in which those two forces “constitute the essence [Wesen] of all chemical as of all mechanical actions” (1812b, 208 bzw. 129). Elsewhere he expressed his desire to demonstrate that the same two fundamental forces of nature manifested in chemical, electrical, and magnetic phenomena are also the “forces that are the cause of impenetrability” (1813, pp. 8–9 [quote on 9]). Ørsted 1812b, p. 65 = 1920a, 2, p. 66.

ØRSTED’S PRESENTATION OF OTHERS’

315

[Geschöpf] that finally stands there in boundless manifoldness [Mannigfaltigkeit] and at the same time contemplative unity. From this point, however, she begins to destroy it again in order to represent the eternal activity and the eternal law in new creatures”—etc.211 He elaborated an analogy between the serial development of plants, animals, and metals in terms of the “developmental stages [Entwicklungsstufen]” they passed through, whereby “all substances in the earth are nothing but resting places of the activity with which nature proceeded from work to work in the formation of the earth.”212 Although the French version kept the basic idea of analogous developmental series of organisms and metals—whereby “in order to follow a course [marche] conformable to nature’s, one must arrange [inorganic] bodies in series in accordance with certain laws”—it made the point much more compactly and eliminated all the characteristically naturphilosophisch terminology.213 Similarly, although the French version kept the original’s language of opposite forces, it had no equivalent for the German’s Gegensätze.214 Nor, in its otherwise very similar treatment of the heat produced by the passage of electricity through less-than-perfect conductors, did the Recherches have language equivalent to the Ansicht’s pronouncement that heat can be regarded “as an internal conflict [Wechselkampf] of the opposite forces.”215 Nor, in presenting his views on the propagation of light via dynamical undulations, did he repeat in French his original comment that “[t]he possibility of such a view has already been recognized by Schelling in his Weltseele.”216 And there is nothing in French to match the language in which he presented the relationship of heat and light to combustion in cases where the unification of Brennkraft and Zündkraft takes place in an undisturbed manner undetectable to our senses: In jener reinen Vereinigung der Kräfte ist also Licht, Wärme, Verbrennung, aber in einer unsern Sinnen unzugänglichen Durchsichtigkeit und Ungetrenntheit. Das Insichseyn ist hier die eigentliche Form des Lichts. Das wechselseitige Ineinandergreifen ist die Form der Wärme, das Identificiren der Thätigkeiten die der Verbrennung.217 One is not surprised that Ørsted’s translator, Marcel de Serres (1783–1862), felt called to rebut the charge that the facility with which the German language forms new words encouraged the Germans to overindulge in abstract ideas and metaphysics.218 Ørsted’s most extensive purging of naturphilosophisch elements from the book was his complete elimination of the original’s section devoted to “General

211

212

213

214 215 216 217 218

Ørsted 1812b, pp. 65–66 = 1920a, 2, p. 66. Given the tenor of the passage, it seemed appropriate to translate the pronouns pertaining to the grammatically feminine Natur as “she” and “her.” Ørsted 1812b, pp. 68–70 (quote on 68) = 1920a, 2, p. 67. Ørsted’s tone in this section is remarkably Lamarckian, though I know of no evidence that he’d read Lamarck. Ørsted 1813, pp. 66–70 (quote on 69). “La marche de la nature” was one of Lamarck’s standard phrases. Cf. Ørsted 1812b, p. 115 = 1920a, 2, p. 88 with the corresponding passages in 1813, pp. 109–110. Ørsted 1812b, p. 165 = 1920a, 2, p. 111; cf. 1813, 155–157. Ørsted 1812b, p. 223 = 1920a, 2, p. 136; cf. 1813, 201–203. Ørsted 1812b, p. 283 = 1920a, 2, p. 162. Serres, in Ørsted 1813, xiv–xv.

316

K. L. CANEVA

Considerations on the Two Fundamental Forces.” In part a synopsis of the doctrine of forces that had gone before, its wording was especially redolent of the tropes of Naturphilosophie. The forces of combustibility and comburence reappeared as different forms of action of the two fundamental forces of nature, which were themselves responsible for the filling of space that (implicitly) we perceive as bodies.219 Now, however, those fundamental forces made their appearance as a singular entity, as “two opposite views of the single space-filling activity,” as manifestations of a single fundamental force (Urkraft), “of which our two forces [are] only different forms of action” or “forms of activity [Thätigkeitsformen].”220 What had been the occasional Gegensatz referred to earlier in the book were now ever new Gegensätze contained in an (ontologically preceding) Gegensatz, producing a “differing depth of the Gegensätze” that extended without predetermined end.221 Here he explicitly identified the lawfulness of nature with reason (Vernunft), asserting that “the laws of nature are identical to the laws of reason”: both are thoughts or ideas such that the essence of the universe is “the epitome of all ideas, identical to absolute reason.”222 He capped this line of thinking with a pronouncement wholly unlike anything his French readers would be exposed to: “And so we see all of nature united in one as the manifestation [Erscheinung] of an infinite force and an infinite reason, as the revelation of God.”223 The last and often quite different sections of the two books—“General Considerations on the State of Chemical Science [Naturlehre]” in the Ansicht, “Considerations on Theories in the Experimental Sciences, and Summary of the Principles of the Dynamical System” in the Recherches—both dealt with general issues concerning the role of theories in science and the importance of directing one’s attention to the discovery of the laws, not the causes, of phenomena. The essential merit of Franklin’s theory of electricity, for example, lay not in its assumption of some material substrate regarded as the cause of the phenomena but in its formulation of the laws of electrical action.224 Only the German version, however, contained a long and largely favorable discussion of the merits of the phlogiston theory, whose defenders should not simply be dismissed as having backed an erroneous theory based on an imagined substance, but praised for having established “the principal law [Hauptgesetz] … that bodies in combusting lose their combustibility, but that a combusted body can be restored to the state of combustibility by means of one still more combustible…. Thus not the invention [Ersinnung] of phlogiston as a cause of combustion but the discovery of a portion of the laws of combustion constitutes the essence of the phlogistic system.”225 Only the German version sought to 219 220

221 222

223 224 225

Ørsted 1812b, pp. 259–260 = 1920a, 2, p. 150. Ørsted 1812b, pp. 257, 258, 259 = 1920a, 2, pp. 151, 151, 152. The chief characteristic of dynamism is that it “allows the whole with all its parts to develop as different forms of one fundamental force [Urkraft]” (267 bzw. 155). Ørsted 1812b, p. 263 = 1920a, 2, p. 153. Ørsted 1812b, pp. 268–270 (quotes on 269, 270) = 1920a, 2, pp. 156–157 (quotes on pp. 156, 156– 157); cf. 288 bzw. 164–165. These connections are explored in greater detail in Caneva 1998, pp. 100–107. Ørsted 1812b, p. 270 = 1920a, 2, p. 157. Ørsted 1812b, pp. 277–278 = 1920a, 2, p. 160; 1813, pp. 244–245. Ørsted 1812b, pp. 280–281 = 1920a, 2, 161; cf. 1813, p. 256.

ØRSTED’S PRESENTATION OF OTHERS’

317

place alchemy in as favorable light as possible, echoing ideas he had floated five years earlier about the reasonableness of correspondences between metals and planets: “For if the metals have developed along with the earth, the latter along with the other heavenly bodies of our solar system (comets as well as planets and their satellites): how very probable—indeed, one might say certain—is it not that both developments have proceeded according to the same laws, only according to different powers [Potenzen]!”226 Only the German version sought to establish a close parallelism between the properties of the metals and the sequence of their discovery, with the noblest metal, gold, being the first one discovered. In addressing a French audience, Ørsted took care to expunge all such expressions of favor towards ideas he had good reason to expect would elicit hoots of derision. In exchange, in a curious pattern of additions presumably intended to align the author with a more respectable tradition in science, the Recherches repeatedly invoked Newton as having set the methodological standard that all proper science must follow.227 Likewise only the Recherches presented the dynamical theory as being “based on a law drawn from experience.”228 Additional features of the revamped image of science portrayed in the Recherches will be discussed at the end of the next section of this paper. Ørsted had clearly taken great pains to present himself in dress tailored to the anticipated scientific tastes of his German and French readers.

7. ØRSTED’S PRESENTATION OF HIS ELECTROMAGNETIC EXPERIMENTS A week before penning the Latin announcement of his epoch-making discovery, Ørsted published a short preliminary report of his findings in a Danish literary journal.229 He there reported having discovered “a means of producing magnetic actions [or effects (Virkninger)] by means of the electrical forces in their galvanic form.”230 Without yet a clear notion of the circular nature of the action of the connecting wire on the magnetic needle, Ørsted explicitly defended its qualification as “magnetic” until such time as a wholly new class of actions might be discovered. He interpreted the new phenomena as indicating that “the electrical forces produce an alteration in metals in such a way that these act forcefully on the magnetic needle.”231 Groping to make sense of what he was seeing, Ørsted was

226

227 228 229

230 231

Ørsted 1812b, p. 291 = 1920a, 2, p. 166. For a consideration of his previously expressed views, see the discussion of his 1807 paper on the history of chemistry in a later section of this paper. Ørsted 1813, v (Serres’ preface), 71, 79; cf. 1812b, pp. 71, 81 = 1920a, 2, pp. 68–69, 73. Ørsted 1813, p. 200; cf. 1812b, p. 211 = 1920b, 2, p. 129. Ørsted 1820a. Although he, like everyone else, was unaware of the existence of this publication, Roberto Martins’ brilliant reconstruction of Ørsted’s conceptual route to the discovery of electromagnetism is only enhanced by its contents; that enterprise, begun in Martins 1986, culminated in Martins 1999. Ørsted 1820a, p. 447 = Teuber 2002, p. 105. Ørsted 1820a, p. 448 = Teuber 2002, p. 106.

318

K. L. CANEVA

clearly not yet in possession of the sought-for law of the phenomena whose form his experiments had not yet made clear to him. That situation changed quickly after he extended his experiments to include cases where the connecting wire was also under the magnetic needle.232 As this first announcement was appearing in print, Ørsted was composing and sending off the short paper that would secure his place in the history of science. Experimenta circa effectum conflictus electrici in acum magneticam: surely there cannot be another epoch-making scientific paper that so few people (it seems) have actually read in its original wording, either upon its appearance in July 1820 or since. Within a few months after Ørsted had sent copies of his paper—consisting of a single folded sheet of four pages—to an unknown number of individuals and learned societies, it appeared in German, French, English, Italian, and Danish translations, often in more than one place.233 Although editor Johann Salomo Christoph Schweigger (1779–1857) reprinted the Latin original, that was basically a for-the-record exercise: virtually everyone other than the first wave of recipients derived their knowledge of Ørsted’s work from one of the translations.234 One can probably safely generalize from the response of his close friend, Christian Samuel Weiss, who’d received Ørsted’s paper in Berlin on 31 July: “But how could you have let your announcement be written and printed in such atrocious Latin? Why not German and French? or at least in tolerable Latin?”235 As far as I can tell, Ørsted never explained why he chose the route he did. Surely he remembered the problems attending the fact that Winterl had published his Prolusiones ad chemiam saeculi decimi noni in Latin. Surely he knew that in 1820 virtually the only contributions to the physical sciences still published in Latin were dissertations, inaugural addresses, and the few papers published by tradition-bound learned societies like the Societas regia scientiarum gottingensis.236 Surely he knew that the single language of overwhelming international choice—and accessibility— would then have been French, a language he commanded. Besides the matter of language, distributing a privately printed pamphlet was just not the way scientists communicated their work to others: I am aware of no other single example from the surrounding decades. Ørsted’s decision was just plain odd, and I really can not adequately explain it except by supposing that he wanted word of his discovery to get out quickly and simultaneously to a widely dispersed scientific community, and that he perhaps fancied that Latin lent it a classical aura of being for the ages.

232 233

234 235 236

The importance of this step was signaled by Torsten Schlichtkrull in Teuber 2002, p. 186. A convenient place to consult some of these translations—in the languages named—is Ørsted 1920c, which reprints them in facsimile (except for the Danish). Parts of this section derive from Caneva 2001, 2–5. An earlier preliminary announcement in Danish (Ørsted 1820a) attracted no attention, and was probably intended (if necessary) to secure his priority. Its relevance for the development of Ørsted’s understanding of just what it was he’d discovered will be examined in a forthcoming paper. Ørsted 1820c. Ampère (1820, pp. 201–202) cited Ørsted in Latin. Weiss, addition of 4 August to his letter of 2 August 1820 to Ørsted, in Ørsted 1920b, 1, p. 270. In a footnote accompanying his reprinting of Ørsted’s Latin text, Schweigger—whose dissertation was in classics, not science—expressed the wish that more use were made of the language common to scholars of all nations (in Ørsted 1820b, p. 275).

ØRSTED’S PRESENTATION OF OTHERS’

319

The announcement is also an odd mix of mostly pure description and cautious generalization, with a whiff of Naturphilosophie provided by the nonstandard and prominently placed term “electrical conflict.” If he’d wanted to remain noncommital, he might have spoken solely in terms of the “connecting wire,” his filum conjungens, just as the effect in question was not presented initially as being on magnetism, but on the magnetic needle. Even the spareness of Ørsted’s presentation, eschewing almost entirely any broader considerations, is a bit surprising since, in another context—in defense of Winterl’s presentation of his chemical system before it could be experimentally established—he allowed that “a new experiment…only acquires its significance…by its relation to the world of ideas [Tankeverden].”237 In subsequent writings Ørsted did not much use the term conflictus electricus (in any language), and his contemporaries—other than Faraday—seem to have paid little attention to his declarations that the electrical conflict is not confined to the connecting wire, but also takes place in the circumjacent space, spiraling around the conductor.238 Ørsted’s electrical conflict thus quickly became assimilated to the current variously qualified as electric, voltaic, or galvanic—as witnessed by the very title of the contemporary English translation: “Experiments on the Effect of a Current of Electricity on the Magnetic Needle.” Indeed, Ørsted himself soon stopped insisting on the fieldlike spatial extendedness of the electromagnetic action. In a follow-up paper on his “Recent Electromagnetic Experiments,” written in September 1820 and published in German, French, and English versions, Ørsted introduced the adjective thenceforth favored to describe the new phenomena, after which electromagnetism soon came to refer simply to the interaction between electric current and magnet.239 Reflecting his still unsettled conceptualization—and terminology—Ørsted nevertheless there referred to his “first experiments on the magnetic action [Wirkung; action] of the galvanic apparatus,” not the action of the electric conflict, electric current, or connecting wire.240 Kirstine Meyer published a selection of Ørsted’s notes on the many and systematically varied experiments he performed as he tried to make coherent sense of the new phenomena, hence we can appreciate the Experimenta’s distillation of those raw results into a few summary statements (“momenta quædam,” roughly “certain main points”).241 The most important for our purposes are the following: One can similarly conclude from what has been observed that this [electrical] conflict describes circles, for this appears to be the condition without which it would not be possible that the same part of the connecting wire which, when placed below the magnetic pole carries it towards the east, when placed above it drives it towards 237 238

239

240

241

Ørsted 1828, p. 523; quoted in Jacobsen 2000, p. 119. Ørsted 1820b, [2], [4] = 1820c, p. 277, 280 = 1920a, 2, pp. 215, 218. He once spoke of the “flow [cursus] of the electrical forces in the connecting wire” ([4] bzw. 281 and 218). In addition to the papers’ titles, see Ørsted 1820d, 365 (“electro-magnetische Wirkungen”), 367 (“electro-magnetische Versuche”) = 1920a, 2, pp. 219, 220; 1820e, p. 78 (“effets électro-magnétiques”), 79 (“expériences électro-magnétiques”); 1820f, p. 375 (“electromagnetic effect”), 376 (“electromagnetic experiments”). In one place where the German has, irregularly, “magnetisch-electrische Wirkung” (365 bzw. 219) the French has “effet électro-magnétique” (78). Ørsted 1820d, p. 364 = 1920a, 2, p. 219 (“Wirkung”); 1820e, p. 78 (“action”). Ørsted 1820f has no independent authority. Meyer 1920, lxxiii–lxxxviii; Ørsted 1820b, [3] = 1820c, p. 280 = 1920a, 2, p. 217.

320

K. L. CANEVA the west, it being the nature of a circle that motions in opposing parts have the opposite direction… All the effects on the north pole set forth here can be easily understood by supposing that the negatively electrical force or matter runs through a spiral line bent towards the right and drives before it the north pole while not acting on the south pole. The effects on the south pole are similarly explained if we attribute to the positively electrical force or matter a contrary motion and the ability to act on the south pole but not also on the north pole. The agreement of this law [lex] with nature will be better perceived by a repetition of the experiments than by a long explanation.242

It is this circular form of the electromagnetic action and this law that Ørsted saw as his chief discoveries. In an account of his electromagnetic researches written in 1827 he characterized his principal discovery as “the fundamental law of electromagnetism, namely that the magnetic effect of the electrical current has a circular motion round it.”243 In recounting his discovery of “electricity’s magnetic action” in his 1828 autobiography, he noted that his first observations were too weak and irregular to allow for the discovery of a “definite law” of the phenomena.244 Renewing his experiments in July with a larger battery, he obtained a much greater effect, so that, after many days’ experimenting, he was able “to find the law governing the effects [or actions (Virkningen)]. As soon as he had found it he hastened to announce his work.”245 We shall return to consider the significance of this concern with establishing the law of the phenomena. From early in his career Ørsted had struggled to attain a systematic understanding of the relationship among magnetism, electricity, the so-called chemical process, heat, and light, especially in terms of their differing forms of action. Following Schelling, in 1805 he speculated that magnetism, electricity, and the chemical process correspond to the three dimensions of space.246 More fruitfully for him, in 1812 he had begun to speak of the different forms of action (Wirkungsformen) or forms of activity (Thätigkeitsformen) in which the opposing fundamental forces of nature manifest themselves.247 It was the pursuit of these speculations that led him to his discovery of the interaction between a magnetic needle and the connecting wire of a galvanic circuit. In a Danish-language discussion of that work he interpreted the new relationships by invoking an echo of his notion of form of action: What we here a moment ago called electricity is not so in the word’s stricter meaning; for the force that in the open galvanic or electric circuit acted in a distinctive manner—under a different form—that we call the electric or galvanic, acts here under an entirely different form that we most appropriately call the magnetic; meanwhile, since magnetism acts under the form of a straight line—that is, the opposing forces separate themselves in precisely opposite directions—[while] the forces here, on the contrary, flow incessantly into each other and form a circular course [Kredslöb], the author has called the action dealt with here electromagnetism.248

242 243 244 245 246 247

248

Ørsted 1820b, [4] = 1820c, 280–281 = 1920a, 2, p. 218. Ørsted 1830, p. 575 = 1920a, 2, p. 358. Ørsted 1828, pp. 536, 537. Ørsted 1828, p. 537. Ørsted 1805a, pp. 18–19 = 1920a, 3, pp. 103–104. From among the many occurrences of Wirkungsform, see Ørsted 1812b, pp. 5, 236, 248, 252, 258 = 1920a, 2, p. 38, 142, 147, 149, 151; for a rarer occurrence of Thätigkeitsform, see 259 bzw. 152. Ørsted 1821a, p. 14 = 1920a, 2, p. 448.

ØRSTED’S PRESENTATION OF OTHERS’

321

Thus it appears here that for Ørsted the principal need for a new term stemmed from the unprecedented circular form of the electromagnetic action and not so much from the fact that it represented an interaction between electricity and magnetism.249 The appreciably expanded German and French essays containing Ørsted’s early “Reflections on Electromagnetism” also characterized it in terms of its peculiar “form of action.”250 In those essays Ørsted further referred to electromagnetism as a peculiar state of the galvanic connecting wire.251 He reasoned as follows: Since new electricity is ceaselessly developed in the galvanic pile, the discharges must be regarded as a ceaselessly renewed give and take. The peculiar state of the forces occurring in the discharging conductor, in which they act as electromagnetic forces, thus seems to me to be a state of ceaseless motion [ein unaufhörlich bewegter]. In the magnet, however, these same forces appear to deviate from the electromagnetic form of action only insofar as they are found in an almost resting state, and form no closed circle.252

Against the theory of transversal magnetism of Austrian chemist Johann Joseph Prechtl (1778–1854), which explained electromagnetic action in terms of the placement of end-to-end loops of small magnets around the circumference of the connecting wire, Ørsted said he “preferred to retain the name electromagnetism for the state of the electrically traversed conductor” because nowhere in such a conductor does a magnetic pole appear and because “the continual production of new electricity in the galvanic apparatus requires that we also assume a renewed electromagnetism, an uninterrupted circular course [Kreislauf] of electrical forces in the conductor under the magnetic form of action. Only where the circular course is interrupted without the cessation of the opposing directions of activity does actual magnetism arise.”253 He retained the language of “form of action [Wirkungsform]” and the image of “the spiral motion of the electromagnetic action” in the last of his contributions to the exposition and explication of the new phenomena, citing his original Latin wording in support of his claim that “electromagnetism forms a circular course (hunc conflictum gyros peragere).”254 If Ørsted’s central interest was indeed in identifying the form or mode of action of the newly detected electromagnetic state in which the universal opposing forces of nature manifest their activity in and around the connecting wire of a galvanic circuit, and in assimilating this form of action to the (never fully formalized) conceptual scheme he had worked to develop over the previous two decades, then we must reconsider what we mean by saying that ‘Ørsted discovered electromagnetism.’ He did indeed discover that a magnetic needle is deflected by some action present in the connecting wire, but that was not the meaning of the discovery for him. While he applied the term ‘electromagnetism’ to his particular

249

250 251 252 253 254

Recall that in the preliminary Danish announcement of his discovery Ørsted defended its qualification as “magnetic” until such time as a wholly new class of actions might be discovered (Ørsted 1820a, pp. 447, 448 = Teuber 2002, pp. 105, 106). Ørsted 1821b, p. 205 = 1920a, 2, p. 227; 1821c, p. 162 = 1821d, p. 3. Ørsted 1821c, p. 165 = 1821d, p. 8. This passage is not in 1821b. Ørsted 1821b, pp. 219–220 = 1920a, 2, 237; 1821c, p. 17&3tilde; 1821d, p. 19. Ørsted 1821b, p. 211 = 1920a, 2, p. 232; 1821c, p. 16&9tilde; 1821d, p. 14. Ørsted 1821e, pp. 127, 125, 124 = 1920a, 2, pp. 248, 247, 246, respectively.

322

K. L. CANEVA

theory of the form of action of spiraling electromagnetic activity, others applied it simply to the new phenomena. And that is the meaning that has come down to us, sanctioned by long-standing consensus. Even in the short run Ørsted did not control the general understanding of just what it was his discovery most significantly consisted in. His presentation of his own work was neither well enough developed—especially with regard to his disinclination to supplement his essentially pictorial representation with any kind of mathematical analysis—nor sufficiently in tune with an increasingly dominant understanding of the proper mode of scientific explanation to allow his interpretation of the phenomena to establish itself as the norm, especially when one considers that that interpretation—insofar as anyone rightly understood it on its own terms—was becoming increasingly unacceptable to an increasing number of scientists. Whereas the discovery of electromagnetism opened up vast new fields of research and analysis for countless others, for Ørsted that work for the most part marked a tacit end to the program of understanding he’d been actively pursuing for twenty years. It is telling in this regard that Ørsted apparently made no attempt to exploit Seebeck’s discovery of (in Ørsted’s terms) thermoelectricity—a discovery he first made known to a wider scientific public—in order to exemplify or extend the ideas on force he had been playing with for decades. Establishing the unprecedented circular nature of the new electromagnetic form of action of the underlying opposing fundamental forces of nature was only one of Ørsted’s original concerns, one closely connected with the enunciation of the law of that action. He gave the matter further consideration in the second section of his German and French “Reflections on Electromagnetism” of mid-to-late 1821, headed “Explanation of the first law of electromagnetic effects,” where he enunciated a revised version of his original law, now provisionally termed a “rule”: The rule I established for the electromagnetic action [or effect (Wirkung)] is—somewhat more developed here—the following: In the meeting of the opposite electrical forces* associated with resistance the former take on a new mode of action [Wirkungsart; forme d’action], as a consequence of which the positive electrical force repels the south end of the magnetic needle [and] attracts the north end, while the negative force repels the north end of the needle [and] attracts the south end**; however, the path of the forces in this is not a straight line but a left-handed spiral or screw-shaped line.

… * ) I repeat here what I have already explained in other works, that by electrical forces I understand nothing but the unknown cause of the electrical phenomena, whether it be bound to an independent material substance or be an autonomous activity. **

) In my first announcement I had assumed only repulsions of the electrical forces against the magnetic, but I soon realized that here, out of fear of assuming more than was demanded, I had fallen into an inconsistency. For if the magnetic forces are the same as the electrical, it automatically follows that opposite forces must attract each other just as well as similar must repel each other. I have, therefore, in my lectures for some time made the aforementioned correction to the above law.255

255

Ørsted 1821b, pp. 202, 202–203 = 1920a, 2, pp. 225–226; 1821c, pp. 163–164; 1821d, pp. 6–7. The French versions omit the last sentence of the second footnote. The words translated “out of fear

ØRSTED’S PRESENTATION OF OTHERS’

323

Although some had reportedly taken exception to this assumption of a spiral motion of the electrical forces, Ørsted was determined to demonstrate that that representation suffices not only to explain but also to predict all of the electromagnetic phenomena, something that he believed could not be said of any of the other proposed modes of representation (whereby his principal rival was Ampère’s electrodynamical explanation of the forces acting between current-carrying wires and magnets supposed to be composed of circular electric currents). Once he had done so he would ask the reader to consider whether his “rule” is not then also a “law…according to which the phenomena are ordered in nature itself.”256 Alas, a footnote told the reader that the author was not yet prepared to make good on this intention. Nevertheless, he did go on to criticize Prechtl’s alternative theory of transversal magnetism (as noted above), and in the next section of the paper— headed in German “Explanation of the interaction of galvanic conductors out of the fundamental law [Grundgesetz]”—he argued (rather weakly) that his essentially pictorial theory was superior to Ampère’s electrodynamics.257 In a Danish essay of 1822 on the progress in chemistry since the beginning of the 18th century, Ørsted described his discovery of electromagnetism as having demonstrated “that the same forces of nature by which the electrical actions are produced also produce the magnetic,” and he expressed “the fundamental law [Grundlov] for the magnetic action of every electric current” in the following graphic (if imprecise) terms: “Every point in the electric current has a magnetic polarity such that an observing eye having the incoming positive electricity on its right side will see the north end of a fine magnetic needle suspended over the point turn away from itself.”258 Ironically, what might well have come down as “Ørsted’s law” specifying the reaction of a magnetic needle to the proximity of a currentcarrying wire has instead become known as “Ampère’s rule” after the man who quickly shifted many researchers’ attention from Ørsted’s electromagnetism to his own electrodynamics. But of course Ampère’s description was published in French, in Paris, and in several places, whereas Ørsted’s was published once, in Copenhagen, and in Danish.259

256 257

258 259

of assuming more than was demanded” from Ørsted 1821b, p. 203 (“aus Furcht mehr anzunehmen als eben gefordert wurde”) were, in the French versions, the somewhat more positivistic “de peur de supposer plus que n’exigeoient les phénomènes” (1821c, p. 164 = 1821d, p. 7). In Danish Ørsted had spoken several times in terms of the “law” he’d discovered, but without either stating it explicitly or otherwise calling attention to its methodological significance (1821a, pp. 14–15 = 1920a, 2, pp. 448–449). With respect to the point made in the first footnote, cf. Ørsted 1812b, pp. 114–115, 288 = 1920a, 2, pp. 88, 164–165 and Caneva 1997, pp. 49–50. See the discussion of these aspects of Ørsted’s work in Meyer 1920, cvii–cviii. Ørsted 1821b, p. 204 = 1920a, 2, p. 226; 1821c, pp. 164–165; 1821d, p. 7. Ørsted 1821b, p. 212 = 1920a, 2, p. 232; the corresponding section in the French versions is headed simply “Explanation of the attractions and repulsions that galvanic conductors exert on each other” (1821c, p. 169 = 1821d, p. 14). Ørsted 1822, cited from 1920a, 3, p. 325. Ampère introduced his rule (règle) for determining the direction of deviation of a magnetic needle in a work published in January 1822 coauthored with Jacques Babinet (1794–1872), professor of physics at the Collège Royale de Saint-Louis (Ampère and Babinet 1822a, pp. 26–27 = 1822b, pp. 191–192). He invoked the image of a tiny human observer imagined to lie in the current-carrying wire facing the needle with the current running from feet to head. Although it cannot be my task

324

K. L. CANEVA

It is precisely the essentially pictorial nature of Ørsted’s understanding of electromagnetic action that makes it all the more noteworthy—and strange—that he did not accompany his Latin announcement with a diagram, a deficiency his colleague at Kiel, Christoph Heinrich Pfaff (1773–1852), quickly brought to his attention. Replying soon after receipt of Ørsted’s paper, Pfaff reported being still in the dark over some of Ørsted’s claims, adding that “[p]erhaps many readers will feel the same way, and a small woodcut might perhaps have very easily dispelled the darkness.”260 Pfaff also had trouble with the Latin. Ørsted’s omission of any diagram is still more surprising when one recalls how prominent a role his many sketches played in his own unraveling of the initially far-from-clear experimental situation. One of the things that makes Ørsted’s overriding concern with establishing the law for the new phenomena especially interesting is that it corresponds precisely to the image of science portrayed in his 1813 book, Researches on the Identity of the Chemical and Electrical Forces, in sections absent or different from the German original. At issue in a new section headed “Considerations on theories in the experimental sciences, and summary of the principles of the dynamical system,” was the relationship between facts and theories and the latter’s role and legitimacy, which Ørsted felt it necessary to defend. The best scientists, he said, have shown “that it is necessary to search for facts as the necessary base of a good theory, and for theory as the only means of conceiving nature, which presents us with an infinity of facts, of which we can examine in detail only a very small number.”261 Nor must one forget, he urged, “that the sciences do not present us with facts, but with our observations on the facts”: That which one presents in the sciences as the simplest fact is already highly composed and is only a summary [résumé] of a very great number of reasonings [raisonnemens] on the simple facts; for a fact is always observed on the basis of one or a few individuals, but science does not speak to us of individuals, [rather] it presents us with an observation as [if it had been] made on the basis of the species, constructed out of an infinity of individuals.262

What he had said of the dynamical theory earlier in the book well captured what his attitude would be seven years later towards his electromagnetic law: “The dynamical theory, which is based on a law drawn from experiment [expérience] and which already explains the phenomena better than the two others, will doubtless be increasingly confirmed by new researches.”263 Ørsted’s translator, with whom he worked very closely in producing the French version of his View of the Chemical

260 261 262 263

here to track the transformation of this image into canonized form, I note its presentation—without any special designation—by Claude-Servais-Mathias Pouillet (1790–1868), professor of physics at the Faculté des Sciences and the École Polytechnique, in his popular physics text (1832, 1, pt. 2, pp. 241–242), and its appearance as “Ampère’s rule” in the 1847 English translation of a popular German textbook of physics (Müller 1848, pp. 440–441). Pfaff, letter to Ørsted of 27 July 1820, in Ørsted 1920b, 2, p. 466. Ørsted 1813, p. 240. Ørsted 1813, p. 241 (both quotes). Ørsted 1813, p. 200; cf. 1812b, p. 200 = 1920a, 2, p. 211, where “die dynamische Theorie” is not spoken of as being “fondée sur une loi tirée de l’expérience.”

ØRSTED’S PRESENTATION OF OTHERS’

325

Laws of Nature, had already signaled in the preface that the goal of science is to discover “general laws,” and he recommended the author’s work as exemplifying the true “spirit of the sciences of observation”: Thus one always follows the wrong path when one wants in the first instance to arrive at the causes of the phenomena, for it is only after having recognized their laws that it is possible to deduce the causes from them. This procedure is also the only one that can be confirmed by experiment, and is the one that has been followed by the inventors of good theories who, by determining certain laws [certaines lois], have prepared the way for all our discoveries. This truth appears to have been well felt by the author of this treatise; and in effect his doctrine is in a certain way only a new way of envisioning the facts and of tying them together.264

The theoretical views concerning the relationship among primitive facts, their idealized representation, and the laws that express the experimentally established regularities of the phenomena as presented in the Researches of 1812 correspond closely both to the image of science Ørsted wished to present to the world in the Experimenta of 1820 and to his actual modus opperandi. They would also appear to reflect Ørsted’s understanding of a style of science he identified not only as methodologically optimal, but also as distinctively French. For the rest, one again glimpses here an underlying tension between Ørsted’s deepest personal concerns and what he thought belonged to proper public scientific practice.

8. ØRSTED’S EDITING OF THE 1851 REPRINTINGS OF HIS 1807 “REFLECTIONS ON THE HISTORY OF CHEMISTRY” In 1807 Ørsted published essentially identical Danish and German versions of a lecture he’d given during the winter semester of 1805–1806 entitled “Reflections on the History of Chemistry.” In 1851 the Danish text was reprinted with appreciable changes in his Collected and Posthumous Works, while the German translation (by someone other than Ørsted) was reprinted with essentially the same changes in the German edition he edited of his Spirit in Nature. In the same year a different (or appreciably reworked) translation showing most of the same changes was published in the German edition of his Collected Works.265 The changes Ørsted made were extensive and significant, and all appear to have had the same purpose: to cast a veil over his earlier enthusiasm for things naturphilosophisch or otherwise far-out. Among the latter I include his thorough recasting of a long (two-and-a-halfpage) passage in which he sought to find some support for the old belief in a correspondence between the planets and metals: such an idea was not, he urged, as fantastical and improbable as it might at first blush appear.266 As he asked 264 265 266

Serres, in Ørsted 1813, i, ii, ii (respectively). See the bibliography entries for details. Ørsted 1807a, pp. 15–18 1807b, pp. 204–206 (= 1920a, 1, pp. 322–324). Page length was gauged after the original Danish version.

326

K. L. CANEVA

rhetorically, “why couldn’t certain metals, which are especially accustomed to accompany others, be regarded as their satellites?”267 In two of the reworked versions this section is reduced to about one-fifth of its original length, and its point is much more tentatively advanced. Even at that, he closed the discussion with a remark further distancing him from such notions: “Yet we would not hide the fact that we are here building supposition upon supposition, and do not lay much weight on this fancied possibility.”268 Compare this to the original ending: “It must suffice for us to have seen that some truth did indeed lie behind that era’s errors.”269 The entire discussion was completely omitted from the other German translation.270 He similarly omitted a page-and-a-third-long elaboration of speculative connections between magnetism—the earth’s in particular—and combustibility and heating ability (Varmetendents, Wärmetendenz), and of the possible influence of the magnetism of other heavenly bodies on changes in the weather and on earthly plants and animals. A time will come when chemistry will extend its realm into astronomy just as mechanics has done: One will then be obliged to regard external motion as a product of internal forces, and all of science will terminate in a cosmogeny. You will be even more enraptured by this hope when I have explained to you Ritter’s great discovery that there exist in every operation of nature definite periodical alternations in both the large and the small—a discovery that will instruct us to cast many a prophetic glance towards the vanished past and the unseen future.271

None of this—certainly not Ritter’s exploded ideas on periodicities in nature— was deemed fit for midcentury readers. Among the more recognizably naturphilosophisch ideas he wished to deemphasize was his attachment to dynamical chemistry, and he omitted a page-and-a-halflong passage in which he crafted an analogy between it as the “higher chemistry” and antiphlogistic (i.e. Lavoisian) chemistry as the “elementary chemistry” on the one side and, on the other, post- and pre-Copernican astronomy.272 The older sciences had made still-valuable observations, and had ushered in their newer and more correct successors, but they were both now equally superseded. The new chemistry (as of 1807) was to be one concerned not with elemental composition but with forces and their various forms: how could one ever know whether one had correctly identified the really elemental fundamental substances, or when one had tallied them all? “If, on the contrary, everything depends on certain fundamental forces and on the forms in which they manifest themselves, then one must

267 268 269 270 271

272

Ørsted 1807a, p. 18 1807b, pp. 205–206 (= 1920a, 1, p. 324). Ørsted 1851–1852, 5 (1851), p. 13 1850–1851a, 2 (1851), p. 391. Ørsted 1807a, p. 18 1807b, 206 (= 1920a, 1, p. 324). Cf. Ørsted 1850–1851b, 3 (1851), p. 154. Ørsted 1807a, p. 31 (which has a paragraph break after “Kosmogonie”)∼ 1807b, p. 215 (= 1920a, 1, p. 331); for the later omissions cf. 1851–1852, 5 (1851), p. 20; 1850–1851a, 2 (1851), p. 406; 1850–1851b, 3 (1851), pp. 161. Ørsted 1807a, pp. 34–35∼ 1807b, pp. 217–218 (words not italicized) (= 1920a, 1, pp. 332–333); for the later omissions cf. 1851–1852, 5 (1851), 21; 1850–1851a, 2 (1851), p. 408; 1850–1851b, 3 (1851), p. 163.

ØRSTED’S PRESENTATION OF OTHERS’

327

be able to discover the principle of these forms and to indicate which and how many are possible, more or less after the example that Schelling has provided us, whereby he represented them in terms of the three dimensions in space.”273 By 1851 Ørsted was no longer eager to associate himself with Schelling’s bythen discredited system. He likewise eliminated the only passage in the lecture— amounting to three-quarters of a page—that mentioned Winterl, despite the fact that, although generally in praise of Winterl’s undertaking, it recognized that Winterl hadn’t provided his otherwise promising system with an adequate experimental base. Perhaps what tainted the passage were Ørsted’s opening and closing sentences, which suggested that Winterl’s day might yet come: “I therefore very much fear that Winterl will not profoundly affect his time period. …Nor does it depend on one man, but on the time period, whether his great work will find the desired testing and purification.”274 (One wonders whether Ørsted intended an allusion to the “great work” of the alchemists.) And perhaps Ørsted did not wish to awaken dormant memories of his own fascination with Winterl’s long-discredited system. His revised history of chemistry was at the same time a tacit recasting of his own scientific image.

9. ASSESSMENT AND CONCLUSIONS One of the principal threads running through the episodes presented here has been Ørsted’s complex relationship to Naturphilosophie—or dynamical science, or whatever else one wishes to call the congeries of like-spirited understandings stemming from the forays into science of Kant’s and Schelling’s philosophies of nature. Ørsted’s quick enthusiasm for the work of Winterl and Ritter reflected his conviction that each constituted a major experimentally grounded step in the direction of a philosophically right-headed Kraftlære (or “chemistry,” as he understood the word in his idiosyncratically expanded fashion). What was announced as an exposition of Winterl’s chemistry thus became a subtle rephrasing of it in language more agreeable to Ørsted’s dynamism—a rephrasing clearly visible in the progression from Winterl’s Prolusiones to Ørsted’s Materialien and thence to his postscript “Letter” to Manthey and his essay in Europa—and it was that Winterl that others largely allowed to stand in for the original. In the process, Ørsted became a major shaper of the largely German dynamical science. When

273

274

Ørsted 1807a, p. 33 (which has “Shelling”)∼ 1807b, 216 (= 1920a, 1, p. 332); for the later omissions cf. 1851–1852, 5 (1851), p. 20; 1850–1851a, 2 (1851), p. 407; 1850–1851b, 3 (1851), p. 162. See also in this regard a new paragraph, interpolated (in square brackets) after a discussion of the explanation of acidity and alkalinity in terms of the fundamental forces, in which neutralization was glossed as a kind of indifferentiation: 1851–1852, 5 (1851), 18 1850–1851a, 2 (1851), pp. 402–403; omitted from 1850–1851b, 3 (1851), p. 160. The Danish reprinting kept the word Indifferentieringer while the German kept Indifferenzirungen. Those paragraphs were interpolated into the original text at 1807a, p. 28 and 1807b, p. 213. Ørsted 1807a, pp. 45–46 (“at Winterl ikke vil gribe meget ind i sin Tidsalder”)∼ 1807b, p. 225 (“daß Winterl nicht tief in sein Zeitalter eingreifen werde”) (= 1920a, 1, p. 339); for the later omissions cf. 1851–1852, 5 (1851), p. 27; 1850–1851a, 2 (1851), p. 419; 1851–1852b, 3 (1851), p. 169.

328

K. L. CANEVA

he wrote for a French audience, Ørsted let his enthusiasm for Ritter’s philosophical leanings be tempered by his sensitivity to the self-consciously less speculative, more positivistic attitude that dominated the spirit of science in France. Similar considerations lay behind the ways he did and did not present his electromagnetic discoveries and theories to the international scientific public. A perennial issue for Ørsted was the need to find an acceptable accommodation between his private concerns and what he had reason to believe would be publically acceptable. Seen sociologically, scientific knowledge consists in whatever one can get away with claiming vis-à-vis the relevant scientific community. Ørsted’s presentations of Ritter’s work to a Parisian audience—both personally and in print—not only made selected aspects of it widely known there, they also established Ørsted’s own initial Parisian identity as a scientist. Although he might still be “Ørstedt,” he was no longer “Vicktred.” That identity was not enhanced, however, when both Ritter’s and Winterl’s work was found to be frequently in error—worse, to be in error because driven by misguided German philosophical speculations. Ørsted thus knew he had to find a delicate balance when, ten years later, he prepared a French edition of the German-language exposition of his chemical views. As we’ve seen, an important part of that recasting consisted in his suppression or playing down of many of the more characteristically naturphilosophisch aspects of his views, at the same time as he still wished to advance an electrical and chemical theory whose inspirational roots clearly lay in German dynamism. Toward the end of his life, in reediting earlier work, Ørsted took more aggressive steps to expunge from the public record his youthful enthusiasm for anything that smacked of Naturphilosophie. Ironically, even in the case of Ørsted’s reporting on Ritter’s work on light— where the expositions he prepared for German readers emphasized much more extensively than did the sparer reports he penned for French readers the wideranging polar analogies and interconnections that gave the work its principal significance for Ritter—Ørsted did not ultimately succeed in defining the terms in which Ritter’s work would come to be understood, canonized eventually as the discovery of ultraviolet light.275 One is thus reminded of the ultimately limited ability of any single person to define the content of science or its language: such success, both short- and long-term, always depends on the interests and conceptual consensuses of the larger scientific community. To paraphrase Ørsted’s judgement with regard to the largely unfavorable reception of Winterl’s chemistry, it does not depend on one man, but on the time period, whether his work will find the desired approval. (Whereby I do not wish to imply that the ultimate viability of any and all ideas in science is simply a function of their agreement with contemporary attitudes: ultimately, well-established findings can change attitudes at least as much as attitudes shape scientists’ initial response to new claims.) Ørsted may well have gotten appropriate credit for his discovery of electromagnetism, but his concept of an “electrical conflict” and his conceptualization of the electromagnetic action in terms of a new circular “form of action” died largely unnoticed. In a similar

275

See the account in Caneva 2001, pp. 11 p. 17, to be expanded in Caneva forthcoming.

ØRSTED’S PRESENTATION OF OTHERS’

329

vein, even if Ørsted’s conceptualization of Seebeck’s discovery as thermoelectricity did ultimately carry the day—driving the rival term thermomagnetism and the rival theories of transversal magnetism into oblivion—that was due not only to his having established an early thermoelectric beachhead in Paris (with Fourier’s assistance), but more lastingly to the fact that scientists in general soon came to regard the phenomena as fundamentally involving the thermal production of an electric current, just as they soon diverted their attention from Seebeck’s Kette, as a device whose operation was still to be understood and measured, to Seebeck’s thermoelectric pile as (simply, as with Ohm) a convenient source of a relatively easily controlled electric current.276 If it is the individual that proposes, it is still the community that disposes. Ørsted’s scientific career itself provides wonderful evidence of the existence and operation of a Europe-wide scientific community during the first half of the 19th century. People traveled; they exchanged visits and letters; they regularly dined or otherwise socialized with each other, often in diverse groups; they performed their pet experiments for each other; they were invited to speak before learned societies. In every case we have looked at, Ørsted’s presentation of others’ work involved in very important ways personal contacts with a large number of people, and his personal connections were an indispensable component of his scientific influence. More publicly, there was widespread translation of a large fraction of those scientific papers regarded as being of some scientific significance. Some scientists— with Ørsted surely among the all-time leaders—saw to it that their papers were published simultaneously in a diversity of journals and languages.277 On the other hand, it is precisely the need for such publications in more than one language that reminds one that not all scientists were as multiply connected as Ørsted, and that on another level there were also localized scientific communities defined by place and language that exhibited appreciable variation among themselves. Ørsted felt very acutely, and at first hand, profound differences between the scientific styles of French and German scientists, and the adaptations he made in his publications intended for those diverse audiences provide us with tangible evidence of how those national differences were perceived. Ørsted’s example reminds us of the important point that published science is written with a particular audience in mind, hence the author’s assessment of that audience will affect what he or she says. To a significant extent, what counts as science in any given context reflects authors’ perceptions of what they can get away with claiming. Another issue that emerged as a red thread through much of the work considered here concerns Ørsted’s lifelong attempt to understand the relationship between theory and experiment and to find a balance between them in his own work. Without a doubt it was Winterl’s theoretical views that aroused Ørsted’s 276 277

These developments will be more thoroughly examined in Caneva forthcoming. One example: Ørsted composed his “Reflections on Electromagnetism” in a German essay he sent to Schweigger for publication in the Journal für Chemie und Physik (Ørsted 1821b). He sent French translations of it to the editors of the Journal de Physique and the Bibliothèque Universelle (Ørsted 1821c and 1821d), as well as to Van Mons, whose journal, the Annales Générales des Sciences Physiques, had ceased publication with the issue for June 1821 (letter to Van Mons of August 8, 1821, in Ørsted 1920b, 2, p. 451).

330

K. L. CANEVA

interest. Although he recognized the necessity of subjecting those views to experimental test and of assessing the validity of Winterl’s many empirical claims, he himself never got very far with that testing, and when Winterl’s more startling claims failed to hold up under the scrutiny of chemists in France and Germany, those chemists largely lost any interest they might have had in Winterl’s theory of acid-base reactions. It was similarly the broad-ranging theoretical implications of Ritter’s work that most caught Ørsted’s attention, and it was the analogical connections of Ritter’s findings with regard to the chemical polarity of light which he emphasized in his German and Danish publications, and which he deemphasized in the French in favor of a more soberly descriptive account. We have noted the attention he gave in his Recherches of 1813 to the question of the relationship between the varied empirical particulars that scientists encounter directly, the laws they legitimately abstract therefrom, and the causes—typically conceived as forces—they then speculatively infer. His understanding of that relationship underlay the way in which he presented his all-important electromagnetic work, emphasizing description and the reduction of the phenomena to a compact law while downplaying the theoretical concerns that had lain closest to his heart for 20 years. Seen in that light, his response to Seebeck’s findings indicates just how far he had gone in the same direction a scant three years later. Nowhere did Ørsted seek to exploit the discovery of a new connection between heat, electricity, and magnetism in order to refine his earlier attempts to understand such relationships in terms of the different forms of action of the underlying primitive forces of nature, or indeed in terms of any theoretical construct. Instead, he contented himself with determining the basic phenomena, building a multiunit pile out of simple bimetallic elements, and attaching Seebeck’s findings to his own discovery of electromagnetism. By 1823 it appears that Ørsted, like most of his French colleagues, had come virtually to ‘see’ an electric current in the connecting wire of a voltaic pile, and he betrayed no sense in which for him the characterization of Seebeck’s new phenomena as “thermoelectric” entailed any kind of theoretical interpretation of them. He had thus come full circle with regard to the relative significance he assigned in his own work to theory and experiment, at least with regard to what he was willing to go public with. Having said that, I would register here a new appreciation on my part for the ingeniousness, and even the success, of Ørsted’s earlier attempts to come to terms conceptually with the manifestations of force exhibited by the phenomena of electricity, magnetism, heat, light, and chemical activity. That he did not ultimately succeed does not surprise us. After all, we do not believe that nature is amenable to that kind of conceptual analysis: energy was the concept needed to do the trick, and Ørsted not only ‘failed’ to come up with it, he failed to recognize the significance of the work of his student, Ludvig August Colding (1815–1888), that was tending in that direction.278 But if one does not know how the story came out, then Ørsted’s conceptual gropings are at least intriguing, and they reveal a disposition 278

Caneva 1998, pp. 50–54, 108–111. That episode offers another example of Ørsted as the presenter of someone else’s work.

ØRSTED’S PRESENTATION OF OTHERS’

331

unwilling to rest content with a merely phenomenological—read superficial— registering of the phenomena such as satisfied some of his French colleagues like Biot.279 Unfortunately for his influence and reputation, the additional fact that Ørsted never seems to have put any appreciable effort into quantifying and measuring any force-related concepts severely restricted any role he might have played in fostering a deeper understanding of these issues. University of North Carolina at Greensboro

BIBLIOGRAPHY Entries for each author are listed chronologically roughly by date of publication, as near as could be determined or inferred. A month in parentheses at the end of the entry indicates the nominal date of the issue in which it was published, although its actual appearance might have been some months later. Such information may be lacking where I was unable to consult the journal directly. Since the numbering (or not) of plates sometimes varies within a single volume between roman and arabic numerals, I have been careful to distinguish between (for example) “pl. 2” and “pl. II.” When the plates are unnumbered, the number (Anzahl) of them is given as (for example) “1 pl.” When a journal was known to contemporaries by its editor (e.g. Gilbert’s Annalen), his name is given in square brackets before the title of the journal. Pagination of a few works has been given where deemed useful. The following abbreviations are used: Journal de Physique = Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts; PV = Procès-verbaux des séances de l’Académie tenues depuis la fondation de l’Institut jusqu’au mois d’août 1835. 10 vols. Hendaye (Basse-Pyrénées): Imprimerie de l’Observatoire d’Abbodia, 1910– 1922. Ampère, André-Marie, 1820. “Suite du Mémoire sur l’Action mutuelle entre deux courans électriques, entre un courant électrique et un aimant ou le globe terrestre, et entre deux aimans.” Annales de Chimie et de Physique, 15, 170–218, pl. 1–5 (October); errata on 223. —— , 1823. “Extrait d’une Lettre de M. Ampère à M. Faraday.” Annales de Chimie et de Physique, 22, 389–400, 1 pl. (April). Dated 18 April 1823. Ampère, André-Marie, and Babinet, Jacques, 1822a. Exposé des nouvelles découvertes sur l’électricité et le magnétisme, de MM. Ørsted, Arago, Ampère, H. Davy, Biot, Erman, Schweiger [sic], de la Rive, etc. Paris: Méquignon-Marvis. ii + 91 p. Published in January. —— , 1822b. “Exposé des nouvelles découvertes sur le magnétisme et l’électricité.” In Jean Riffault, ed., Supplément à la traduction française de la cinquième édition du Système de chimie, par Th. Thomson (Paris: Méquignon-Marvis), 168–256, in section “De l’électricité,” 163–256. With different pagination, identical to Ampère and Babinet 1822a without the table of contents on i–ii. Anonymous, 1800. Synopsis of the experimental portion of Winterl’s treatise, “Experimenta et observationes de causa aciditatis.” Göttingische Anzeigen von gelehrten Sachen, 2 [des Jahres], No. 88/89 (2 June), 873–884. —— , 1806. Review of Winterl 1800, 1803, and 1804 under the heading “Chemie.” Allgemeine Literatur-Zeitung (Halle and Leipzig), 1 [des Jahres], Nos. 44 and 45 (20 and 21 February), cols. 345–350 and 353–360. —— , 1823. “Extrait des Séances de l’Académie des Sciences.” Annales de Chimie et de Physique, 22, 315–320 (March). Bernoulli, Christoph, 1804. “Nouvelles découvertes galvaniques; par M. Ritter.” Extraites d’une lettre de M. Cristophe Bernouilli [sic]. [Van Mons’] Journal de Chimie et de Physique, 6, No. 17 (15 Pluviôse, an XII = 5 February 1804), 133–141. Probably written during the second half of July 1804; for dating see note 126 and Bernoulli 1805. —— , 1805. “New Galvanic Discoveries by Mr. Ritter, extracted from a letter from Mr. Christ. Bernoulli.” [Nicholson’s] Journal of Natural Philosophy, Chemistry, and the Arts, 8vo ed. [i.e. 279

Ørsted explained his preferences in this regard in a talk delivered in 1830 at the ninth meeting of the Gesellschaft deutscher Naturforscher und Ärzte in Hamburg (Ørsted 1831).

332

K. L. CANEVA

N.S.], 12, 99–103 (October). In a footnote to the title, the date of the issue containing Bernoulli 1804 is given as March 1805. —— , 1806. “Einiges aus Herrn Christoph Bernouilli’s [sic] Erzählung von neuen galvani’schen Entdeckungen des Herrn Akadem. Ritter, in ausländischen Journalen.” [Gilbert’s] Annalen der Physik, 24, 101–103 (September). Partial translation of Bernoulli 1804. Biot, Jean-Baptiste, 1802. Rapport fait à la classe des sciences mathématiques et physiques de l’Institut national, Sur le prix fondé par le premier Consul, pour les découvertes relatives à l’électricité et au galvanisme. [Paris]: Baudouin, Imprimeur de l’Institut national, [1802]. Signed by Laplace, Hallé, Coulomb, Haüy, and Biot, rapporteur. Dated 11 Messidor, an X = 30 June 1802 and read at the public session of 17 Messidor, an X = July 6, 1802 (cf. PV, 2 [1912], 518 [session of 11 Messidor]). —— , 1803. “Rapport fait à la Classe des Sciences Physiques et Mathématiques de l’Institut National par les Cns Laplace, Coulomb, Hallé, Guyton et Biot, sur les recherches faites pendant l’an XI, relativement au Galvanisme.” PV, 3 (1913), 11–17. Delivered at the session of 24 Vendémiaire, an XII = 17 October 1803. —— , 1804. “Rapport fait à la classe des sciences mathématiques et physiques de l’Institut national, sur le prix fondé par le premier Consul pour les découvertes relatives à l’électricité et au galvanisme.” Mémoires de l’Institut National des Science et Arts. Sciences mathématiques et physiques, 5 (Fructidor, an XII [= August/September 1804]), section “Histoire,” 233–238, in section “Distribution de prix,” 223–238. Reprints the text of Biot 1802. Blaineville, Henri-Maire Ducrotay de, 1823a. “Analyse des principaux travaux dans les Sciences physiques, publiés dans le cours de l’année 1823 [i.e. 1822].” Journal de Physique, 96, v–cxxiv (January). —— , 1823b. “Note sur les premières expériences thermo-électriques de M. Seebeck, et sur celles de MM. Œrsted et Fourier.” Journal de Physique, 96, 199–202 (April). Published anonymously; Blainville was the editor. Brugnatelli, Luigi Valentino, 1804. “Extrait d’une lettre de Brugnatelli. Non-existence de la pile à charger.” [Van Mons’] Journal de Chimie et de Physique, 6, No. 16 (15 Nivôse, an XII = 9 January 1804), 132. —— , 1805. “Nicht-Existenz der so genannten Ladungssäule Ritter’s.” [Gilbert’s] Annalen der Physik, 19, 490–491 (April). Translation of Brugnatelli 1804. —— , 1806a. “Observations chimico-galvaniques.” Journal de Physique, 62, 298–318; 1 pl. (April). —— , 1806b. “Chemisch-galvanische Beobachtungen.” [Gilbert’s] Annalen der Physik, 23, 177–219, pl. IV (June). Translation of Brugnatelli 1806a. Caneva, Kenneth L., 1997. “Physics and Naturphilosophie: A Reconnaissance.” History of Science, 35, 35–106 —— , 1998. “Colding, Ørsted, and the Meanings of Force.” Historical Studies in the Physical and Biological Sciences, 28, Pt. 1, 1997 [published May 1998], 1–138. —— , 2001. The Form and Function of Scientific Discoveries. Dibner Library Lecture, November 16, 2000. Washington, D.C: Smithsonian Institution Libraries. viii + 32 p. —— , forthcoming. “ ‘Discovery’ as a Site for the Collective Construction of Scientific Knowledge.” Chenevix, Richard, 1803–1804. “Remarques sur un ouvrage intitulé: Materialien zu einer Chemie des neunzehnten jahrhunderts, etc., ou Matériaux pour servir de bases à une Chimie du dix-neuvième siècle; par le Docteur H. Ørsted.” [Van Mons’] Journal de Chimie et de Physique, 5, No. 15 (15 Frimaire, an XII = 7 December 1803), 241–256; 6, No. 16 (15 Nivôse, an XII = January 9, 1804), 5–16. —— , 1804. “Remarques sur un ouvrage intitulé Materialien zu einer chemie des neunzehnten iahrhunderts etc. ou Matériaux pour servir de base à une chimie du dix-neuvième siècle, publié par D.-J.-B. Ørsted, Ratisbonne 1803.” Annales de Chimie, 50, 173–194 (Floréal = April/May). Slightly different from Chenevix 1803–1804. —— , 1805. “Bemerkungen über ein Werk, welches den Titlel führt: Materialien zu einer Chemie des neunzehnten Jahrhunderts, herausgegeben vom Dr. J. B. Ørsted, Regensburg 1803.” Annalen der Physik, 20, 422–454 (August). Translation of Chenevix 1804. Contains long footnotes by Gilbert containing long quotations from Winterl, Ritter, and others. Christensen, Dan Charly, 1995. “The Ørsted-Ritter Partnership and the Birth of Romantic Natural Philosophy.” Annals of Science, 52, 153–185. Crosland, Maurice, 1967. The Society of Arcueil: A View of French Science at the Time of Napoleon I. London: Heinemann; Cambridge, MA: Harvard University Press. Döbereiner, Johann Wolfgang, 1823. “Aus einem Schreiben vom Prof. Döbereiner an Gilbert. (Phytochemie; Eschwegit; merkwürdige Veränderdung von Holz durch den Blitz; Wiederholung Seebeck’scher Versuche über magnetische Electromotoren durch Erhitzung.)” [Gilbert’s] Annalen der Physik, 73, 111–116, pl. I (January). Dated 12 January 1823. Euler, Karl-Joachim, 1981. “Von Ritter bis Tudor. Zur Erfindung des Bleiakkumulators.” TechnikGeschichte, 48, 28–46.

ØRSTED’S PRESENTATION OF OTHERS’

333

Fourier, Jean-Baptiste-Joseph, 1826. “Analyse des Travaux de l’Académie royale des Sciences, pendant l’année 1822. Partie mathématique.” Mémoires de l’Académie Royale des Sciences de l’Institut de France. Années 1821 et 1822, 5, “Histoire de l’Académie,” 231–320. Read April 24, 1823. Gehlen, Adolph Ferdinand, 1804. “Winterl’s System betreffend.” [Gehlen’s] Neues allgemeines Journal der Chemie, 3, 105–108 (January or February?). Signed “G.” Gehler, 1825–1845. Johann Samuel Traugott Gehler’s Physikalisches Wörterbuch, neu bearbeitet von Brandes, Gmelin, Horner, Muncke, Pfaff. 11 vols. in 22 pts. Leipzig: E. B. Schuckert. Hansteen, Christopher, 1823. “Mathematisk Theorie over de af Prof. Örsted opdagede elektromagnetiske Phænomener.” Magazin for Naturvidenskaberne, 1 (i.e. “Förste Aargangs förste Bind”), 274– 320, pl. 4; 2 (i.e. “Förste Aargangs andet Bind”), 72–117, pl. 1 (June and September?). The second part has “Særsyn” for “Phænomener.” The first part was written at the end of 1821 (2, 72). Hermbstädt, Sigismund Friedrich, 1804. Review of Ørsted 1803a. Neue allgemeine deutsche Bibliothek, 88, St. 2, Hft. 7, 468–470 (c. April). Published anonymously; authorship attributed in Meyer 1920, xxvii. Humboldt, Alexander von, 1993. Über das Universum: Die Kosmosvorträge 1827/28 in der Berliner Singakademie. Ed. Jürgen Hamel and Klaus-Harro Tiemann. Frankfurt am Main: Insel-Verlag. Sixteen weekly lectures delivered between 6 December 1827 and 27 March 1828. Jacobsen, Anja Skaar, 2000. Between Naturphilosophie and Tradition: Hans Christian Ørsted’s Dynamical Chemistry. Ph.D. dissertation, Department of the History of Science, University of Aarhus. —— , 2001. “Spirit and Unity: Ørsted’s Fascination by Winterl’s Chemistry.” Centaurus, 43, 184–218. —— , 2002. “Ørsted, kemien, Ritter og Winterl.” In Ørsted 2002, 55–68. Authorship confirmed by author. Kleinert, Andreas, 1984. “Die Entdeckung der unsichtbaren Strahlen des Sonnenspektrums.” Gesnerus, 41, 291–298. Marum, Martinus van, 1803a. “Expériences avec les nouveaux appareils galvaniques de Ritter; et imitation de quelques-unes de ces expériences par l’appareil électrique ordinaire.” [Van Mons’] Journal de Chimie et de Physique, 5, No. 14 (15 Brumaire, an XII = 7 November 1803), 212–215. Footnote to authors’ names (van Marum and Ørsted) says “Extraites d’une Lettre de M. Van Marum, au Rédacteur.” —— , 1803b. “Lettre de van Marum à J. C. Delamétherie, Sur les expériences galvaniques de Ritter.” Journal de Physique, 57, 471–473 (Frimaire, an XII = November/December 1803). —— , 1804. “Letter from Van Marum to J. C. Delamétherie, on Ritter’s Galvanic Experiments.” [Nicholson’s] Journal of Natural Philosophy, Chemistry, and the Arts, 8vo ed. [i.e. N.S.], 8, 212–214 (July). Translation of Marum 1803b. —— , 1805. “Einige Galvani’sche und electrische Versuche, angestellt im Teyler’schen Museum zu Haarlem.” [Gilbert’s] Annalen der Physik, 19, 488–490 (April). Translation of Marum 1803a. Meyer, Kirstine, 1920. “The Scientific Life and Works of H. C. Ørsted.” In Ørsted 1920a, 1, xi–clxvi. Mons, Jean-Baptiste van, 1803. “Invention d’une pile de charge ou batterie galvanique; et isolément des fonctions de la pile.” [Van Mons’] Journal de Chimie et de Physique, 5, No. 14 (15 Brumaire, an XII = 7 November 1803), 200–206. Published anonymously; authorship inferred from editorship. Müller, Johann Heinrich Jacob, 1848. Principles of Physics and Meteorology. First American edition. Philadelphia: Lea & Blanchard. Originally published London: H. Bailliere [etc.], 1847 (not seen). Author identified as “J. Muller” on title page. Muncke, Georg Wilhelm, 1836. “Magnetismus.” In Gehler 1825–1845, 6, Abth. 2, 639–1162. Nielsen, Hans Toftlund, 2000. “H. C. Ørsteds ‘Chemie.’ ” Dansk Kemi, 81, no. 3, 27–30. Nielsen, Keld, 1989–1991. “Another Kind of Light: The Work of T. J. Seebeck and His Collaboration with Goethe.” Parts I and II. Historical Studies in the Physical and Biological Sciences, 20 (1989–1990), Pt. 1, 107–178; 21 (1990–1991), Pt. 2, 317–397. Ørsted, Hans Christian, 1803a. Materialien zu einer Chemie des neunzehnten Jahrhunderts. Erstes Stück. Regensburg: Montag- und Weißische Buchhandlung. xxvi + 152 p. Preface dated Jena, August 1802. Reprinted in, and cited from, Ørsted 1920a, 1, 133–210; translated in Ørsted 1998, 120–165. —— , 1803b. “Expériences sur les rayons invisibles du spectre solaire. (Note communiquée par M. Vicktred, docteur à l’université de Copenhague.)” Bulletin des Sciences, par la Société philomatique, 7e année, No. 1 (= No. 73, an XI), 197–198 (Germinal = March/April). Signed “I. B.,” presumably Jean-Baptiste Biot. For authorship see note 74. —— , 1803c. “Uebersicht der neuesten Fortschritte der Physik.” Europa. Eine Zeitschrift, 1, Hft. 2, 20–48 (c. June). Signed “O.” Reprinted in Ørsted 1920a, 1, 112–131; translated in Ørsted 1998, 107–119. —— , 1803d. “Experiments on the Invisible Rays of the Solar Spectrum.” [Nicholson’s] Journal of Natural Philosophy, Chemistry, and the Arts, 8vo ed. [i.e. N.S.], 5, 255–256 (August). Translation of Ørsted 1803b.

334

K. L. CANEVA

—— , 1803e. “Expériences sur un Appareil à Charger d’Électricité par la Colonne Électrique de Volta; par M. Ritter, à Jena; présentées à l’Institut National par J.-C. Ørsted, Docteur à l’Université de Copenhague.” Journal du Galvanisme; de Vaccine, etc., 2 (an XII/1803), 97–123, 145–158 (October and November?). To the original memoir (97–123, 145–156) is added a “Post-Scriptum” (156– 158). Presented 28 Thermidor, an XI = 16 August 1803 (PV, 2 [1912], 692). —— , 1803f. “Expériences sur un appareil à charger d’électricité par la colonne électrique de Volta. Par M. Ritter, à Jena. Présentées à l’Institut national par J. C. Ørsted, docteur à l’Université de Copenhague.” Journal de Physique, 57, 345–368 (Brumaire, an XII = October/November 1803). Reprinted in Ørsted 1920a, 1, 214–237. To the original memoir (345–363 bzw. 214–232) are added a “Post-scriptum” (363–364 bzw. 232–233) and an “Addition au mémoire sur la pile à charger de Ritter” (364–368 bzw. 233–237). —— , 1803g. “Expériences sur la lumière; par M. Ritter, à Jena, communiquées par Ørsted, docteur à l’université de Copenhague.” Journal de Physique, 57, 409–411 (Frimaire, an XII = November/ December 1803). Reprinted in Ørsted 1920a, 1, 245–248. Probably read at the Société Philomatique on 18 March 1803 = 27 Ventôse, an XI (Ørsted 1803g, 410 = 1920a, 1, 247; 1870, 1, 125). —— 1803h. “Expériences avec la pile électrique, faites par M. Ritter, à Jena; communiquées par M. Ørsted.” Journal de Physique, 57, 401–405 (Frimaire, an XII = November/December 1803). Reprinted in Ørsted 1920a, 1, 237–242. —— , 1803i. “Expériences sur le magnétisme, par M. Ritter, à Jena; communiquées par Ørsted, docteur ä l’Université de Copenhague.” Journal de Physique, 57, 406–409 (Frimaire, an XII = November/ December 1803). Reprinted in Ørsted 1920a, 1, 242–245. —— , 1804a. “Experiments on Light; by Mr. Ritter, of Jena. Communicated by Dr. Ørsted.” [Nicholson’s] Journal of Natural Philosophy, Chemistry, and the Arts, 8vo ed. [i.e. N. S.], 8, 214–216 (July). Translation of Ørsted 1803g. —— , 1804b. “[Ueber verschiedene physikalisch-chemische Gegenstände].” [Gehlen’s] Neues allgemeines Journal der Chemie, 3, 322–324, in the section headed “Correspondenz” (September or October?). Reprinted under the title “Correspondenz” in Ørsted 1920a, 3, 96–105 and as a letter in Ørsted 1920b, 2, 356–358; translated in Ørsted 1998, 166–167. Dated 10 September 1804. —— , 1804c. “Indbydelse til physiske og chemiske Forelæsninger.” Kjøbenhavnske lærde Efterretninger for Aar 1804, No. 39, 619–621 (ca. October). Reprinted in Ørsted 1920, 3, 78–79. —— , 1805a. “Om Overensstemmelsen mellem de elektriske Figurer og de organiske Former.” Det skandinaviske Litteraturselskabs Skrifter, 1 (i.e. “Förste Aargangs Förste Bind”), 1–22. Reprinted in Ørsted 1920a, 3, 96–105; translated in Ørsted 1998, 185–191. —— , 1805b. “Kritik over den saakaldede Eudiometrie, med Hensyn paa Lægekunsten.” Nyt Bibliothek for Physik, Medicin og Oeconomie, 8, 52–79. Reprinted in Ørsted 1920a, 1, 248–261; translated in Ørsted 1998, 170–179. Read at the Kongelige Medicinske Selskab in January 1805. —— , 1805c. “Kritik der sogenanneten Eudiometrie.” [Gehlen’s] Neues allgemeines Journal der Chemie, 5, 365–392 (October?). Not simply a translation of 1805b. —— , 1805d. “Nye Undersögelser over det Spörgsmaal: Hvad er Chemie?” Det skandinaviske Litteraturselskabs Skrifter, 2 (i.e., “Förste Aargangs Andet Bind”), 240–262. Reprinted in Ørsted 1920a, 3, 105–116; translated in Ørsted 1998, 192–199. —— , 1806a. “[Gegen Volta’s chemische Erklärung der Ladungssäule und über Jürgensen’s Thermometer].” [Gehlen’s] Neues allgemeines Journal der Chemie, 6, 500–502, pl. II, in section “Correspondenz.” Reprinted under the title “Correspondenz” in Ørsted 1920a, 1, 273–276. Dated 4 March 1806; translated in Ørsted 1998, 215–217. —— , 1806b. “Versuche, veranlaßt durch einige Stellen in Winterl’s Schriften.” [Gehlen’s] Journal für die Chemie und Physik, 1, 276–294 (August). Reprinted in Ørsted 1920a, 1, 277–289; translated in Ørsted 1998, 218–226. —— , 1806c. “Die Reihe der Säuren und Basen.” [Gehlen’s] Journal für die Chemie und Physik, 2, 509–547 (December). Reprinted in Ørsted 1920a, 1, 289–315; translated in Ørsted 1998, 227–242. —— , 1807a. “Betragtninger over Chemiens Historie, en Forelæsning.” Det skandinaviske Litteraturselskabs Skrifter, 2 (i.e. “3. Aargangs 2. Bind”), 1–54. Reprinted with appreciable editing in Ørsted 1851–1852, 5 (1851), 1–32. Translated in Ørsted 1998, 243–260. From a series of lectures delivered during the winter semester of 1805–1806. —— , 1807b. “Betrachtungen über die Geschichte der Chemie; eine Vorlesung.” [Gehlen’s] Journal für die Chemie und Physik, 3, 194–231 (February). Reprinted in Ørsted 1920a, 1, 315–343. Reprinted with appreciable editing in Ørsted 1850–1851a, 2 (1851), 371–428; a different (or reworked) translation showing the same appreciable omissions published in Ørsted 1850–1851b, 3 (1851), 143–174. Translated by someone other than Ørsted from 1807a. —— , 1812a. “Kraftlæren.” Printed but never published pages (pp. 1–224, without a title page), probably intended as the second volume to his Videnskaben om Naturens almindelige Love, the first

ØRSTED’S PRESENTATION OF OTHERS’

335

volume of which was published in 1809. For identification and dating see H. Nielsen, 2000, 28–29 and Jacobsen 2000, 125–129, 256. I am indebted to Anja Skaar Jacobsen for a photocopy. —— , 1812b. Ansicht der chemischen Naturgesetze, durch die neueren Entdeckungen gewonnen. Berlin: In der Realschulbuchhandlung. 298 p, 1 pl. Reprinted in Ørsted 1920a, 2, 35–169; translated in Ørsted 1998, 310–392. —— , 1813. Recherches sur l’identité des forces chimiques et électriques. Trans. by Marcel de Serres. Paris: J.-G. Dentu. The “Post-Scriptum” to the “Introduction,” pp. 12–20, reprinted in Ørsted 1920a, 2, 173–177. —— , 1816. “[Theorie over Lyset.]” Oversigt over det Kongelige Danske Videnskabernes Selskabs Forhandlinger og dets Medlemmers Arbeider fra 1 November 1815 til 31 Mai 1816, 12–15. Reprinted in Ørsted 1920a, 2, 433–435; translated in Ørsted 1998, 397–399. Read 1 March 1816. —— , 1820a. “[Middel til at frembringe magnetiske Virkninger formedelst de electriske Kræfter].” Dansk Litteratur-Tidende for Aaret 1820, No. 28, 447–448 (ca. mid-July), in section “Insendt.” Reprinted in Teuber 2002, 105–106 (plus a facsimile on 102 and 104). For dating see also Torsten Schlichtkrull’s note in Teuber 2002, 184–186, where he tentatively dated its composition to 15 July. —— , 1820b. Experimenta circa effectum conflictus electrici in acum magneticam. Hafniæ: Typis Schultzianis. 4 p. Reprinted in Ørsted 1920a, 2, 214–218; translated in Ørsted 1998, 413–416. Dated 21 July 1820. —— , 1820c. “Experimenta circa effectum Conflictus electrici in Acum magneticam.” [Schweigger’s] Journal für Chemie und Physik, 29, 275–281 (July). Reprints Ørsted 1820b. —— , 1821a. “[Undersøgelser over Magnetismen].” Oversigt over det Kongelige Danske Videnskabernes Selskabs Forhandlinger og dets Medlemmers Arbeider fra Mai 1820 til Mai 1821, 12–21. Reprinted as “Meddelelse om Electromagnetismens Opdagelse” in Ørsted 1920a, 2, 447–453; translated in Ørsted 1998, 425–429. Based on presentations of 5 January and 6 April 1821, plus later additions. —— , 1821b. “Betrachtungen über den Elektromagnetismus.” [Schweigger’s] Journal für Chemie und Physik, 32 (= Jahrbuch der Chemie und Physik, 2), 199–231, pl. IV (June). Reprinted in Ørsted 1920a, 2, 223–245; translated in Ørsted 1998, 430–445. —— , 1821c. “Considérations sur l’électro-magnétisme.” Journal de Physique, 93, 161–180, 1 pl. (September). Translation of Ørsted 1821b (letter to Van Mons of 8 August 1821 in Ørsted 1920b, 2, 451). —— , 1821d. “Considérations sur l’électro-magnétisme,…addressées aux Rédacteurs de la Bibliothèque Universelle.” Bibliothèque Universelle des Sciences, Belles-Lettres et Arts. Sciences et Arts, 18, 3–29, pl. I (September), plus errata on p. 164. A revised version of Ørsted 1821c. —— , 1821e. “Schreiben des Herrn Professor Ørsted an die Redaction vom 9. Sept. 1821.” [Schweigger’s] Journal für Chemie und Physik, 33 (= Jahrbuch der Chemie und Physik, 3), 123–131 (September). Reprinted in Ørsted 1920a, 2, 246–251; translated in Ørsted 1998, 446–449. —— , 1822. “Udsigt over Chemiens Fremskridt siden det attende Aarhundredes Begyndelse.” Tidsskrift for Naturvidenskaberne, 1, 1–55. Reprinted in Ørsted 1920a, 3, 301–329. —— , 1823a. “[Theorie over Lyset.]” Det Kongelige danske Videnskabernes Selskabs philosophiske og historiske Afhandlinger, 1, section “Oversigt over det Kongelige danske Videnskabernes Selskabs Forhandlinger og det Medlemmers Arbeider fra 1814 til 1822,” xvi–xix. Same as Ørsted 1816. —— , 1823b. “Nouvelles expériences de M. Seebeck sur les actions électro-magnétiques. (Note communiquée par M. Ørsted).” Annales de Chimie et de Physique, 22, 199–201 (February). Reprinted in Ørsted 1920a, 2, 263–265; translated in Ørsted 1998, 462–463. Probably the text of his report to the Académie des Sciences on 3 March 1823 (cf. Anonymous 1823, 315; PV, 7 [1916], 455). —— , 1823c. “Notiz von neuen elektrisch-magnetischen Versuchen des Herrn Seebeck in Berlin.” [Gilbert’s] Annalen der Physik, 73 (= Annalen der Physik und der physikalischen Chemie, 13), 430–432 (April). Gilbert’s translation of Ørsted 1823b. —— , 1823d. “Sur quelques nouvelles expériences thermo-électriques faites par M. le baron Fourier et M. Ørsted. (Notice lue à l’Académie des sciences par M. Ørsted).” Annales de Chimie et Physique, 22, 375–389, pl. [II] (April). Reprinted in Ørsted 1920a, 2, 272–282; translated in Ørsted 1998, 470–477. Read 31 March 1823 (PV, 7 [1916], 466). —— , 1823e. “[Om Seebecks nye thermoelectriske Kiæde].” Oversigt over det Kongelige Danske Videnskabernes Selskabs Forhandlinger fra Mai 1822 til Mai 1823, 9–10. Reprinted in Ørsted 1920a, 2, 461–462. Not read at a meeting of the society. —— , 1823f. “Electro-Magnetism. New Experiments by M. Seebeck on Electro-Magnetic Action.” Quarterly Journal of Science, Literature, and the Arts, 15, No. 30 (July), 374. Published anonymously; a translation of the first two paragraphs of Ørsted 1823b. —— , 1828. “Ørsted (Hans Christian).” In Hans Ancher Kofod, ed., Conversations-Lexicon, eller encyclopædisk Haandbog… oversat efter den tydske Originals sidste Oplag, med adskillige Forandringer

336

K. L. CANEVA

og Tillæg (28 vols.; Copenhagen: A. Soldin, 1816–1828), 28, 515–541. Reprinted in Ørsted 2002, 13–40 (which records the original pagination). —— , 1830. “Thermo-electricity.” The Edinburgh Encycloædia, ed. David Brewster (18 vols.; Edinburgh: William Blackwood), 18, 573–589, pl. DXXII–DXXIII. Reprinted in Ørsted 1920a, 2, 351–398 and in Ørsted 1998, 542–582. Sent to Brewster on 6 July 1827 (Ørsted 1920b, 2, 279–280). Published anonymously. —— , 1831. “[Ü]ber die Verschiedenheit des physikalischen Vortrages von dem mathematischen, auch wenn beyde dieselben Wahrheiten darstellen.” Isis von Oken, [24], cols. 854–857 (August). A close paraphrase (by Oken?) of a talk delivered on 20 September 1830 at the ninth Versammlung deutscher Naturforscher und Ärzte in Hamburg. —— , 1847. Undated autobiographical sketch (in French) prepared for the Biographie des hommes du jour, ed. Germain Sarrut and B. Saint-Edme (i.e. Edme-Théodore Bourg) (7 vols.; Paris: H. Krabbe, 1835–1843 [to cite the Bibliothèque Nationale entry]). Manuscript in Det Kongelige Bibliotek, Copenhagen (Ørsted 4, 2°); 8 unnumbered pages plus 2-page letter from P. H. Krabbe dated January 1847. Although assigned here to 1847, it may in fact have been written later. I am indebted to Anja Skaar Jacobsen for a photocopy. —— , 1850–1851a. Der Geist in der Natur. Deutsche Original-Ausgabe des Verfassers. 2 vols. Munich: Literarisch-artistische Anstalt der J. G. Cotta’schen Buchhandlung. —— , 1850–1851b. Gesammelte Schriften. Trans. Karl Ludwig Kannegiesser. 6 vols. Leipzig: Carl B. Lorck. —— , 1851–1852. Samlede og efterladte Skrifter. 9 vols. Copenhagen: Andr. Fred. Høst. —— , 1870. Breve fra og til Hans Christian Ørsted. Ed. Mathilde Ørsted. 2 vols. Copenhagen: Th. Linds Forlag. —— , 1920a. Naturvidenskabelige Skrifter (= Scientific Papers). Edited by Kirstine Meyer. 3 vols. Copenhagen: Høst. —— , 1920b. Correspondance de H. C. Örsted avec divers savants. Ed. Marius Christian Harding. 2 vols. Copenhagen: H. Aschehoug. —— , 1920c. La Découverte de l’Electromagnétisme faite en 1820 par J.-C. Œrsted. Copenhagen: H. H. Thiels Bogtrykkeri. —— , 1998. Selected Scientific Works of Hans Christian Ørsted. Translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen. Introduced by Andrew D. Wilson. Princeton, NJ: Princeton University Press. —— , 2002. H. C. Ørsteds Selvbiografi. Ed. Anja Skaar Jacobsen and Svend Larsen. [Århus]: Steno Museets Venner. Ostwald, Wilhelm, 1896. Elektrochemie, ihre Geschichte und Lehre. Leipzig: Veit & Comp. Pfaff, Christoph Heinrich, 1838. “Thermoelektricität. Thermomagnetismus.” In Gehler 1825–1845, 9, Abth. 1, 731–823. Poggendorff, Johann Christian, 1841. “Gedächtnißrede auf Thomas Johann Seebeck.” Abhandlungen der Königlichen Akademie der Wissenschaften zu Berlin. Aus dem Jahre 1839, xix–xxxviii. Delivered 4 July 1839. Poppe, Kurt, 1972. “Johann Wilhelm Ritter und Ernst II., Herzog von Sachsen-Gotha etc. Zwei unbekannte Briefe aus den Jahren 1802–1803.” Jahrbuch des freien deutschen Hochstifts, N. S., 177–202. Pouillet, Claude-Servais-Mathias, 1832. Élémens de physique expérimentale et du météorologie. 2nd ed. 2 vols. in 4 pts. Paris: Béchet jeune. Rehm, Else, 1971. “Unbekannte Briefe Johann Wilhelm Ritters an Arnim, Savigny, Frommann, Schelling und andere aus den Jahren 1800–1803.” Jahrbuch des freien deutschen Hochstiftes, N.S., 10, 32–89. Ritter, Johann Wilhelm, 1801a. “Chemische Polarität im Licht. Ein mittelbares Resultat der neuern Untersuchungen über den Galvanismus.” Litteratur-Zeitung (Erlangen), Intelligenzblatt, No. 16 (18 April), cols. 121–123. —— , 1801b. “[Auffindung nicht sichtbarer Sonnenstrahlen außerhalb des Farbenspectrums, an der Seite des Violets.]” [Gilbert’s] Annalen der Physik, 7, 527, in a section headed “Auszüge aus Briefen an den Herausgeber,” 501–528, “6. Von den Herren Ritter und Böckmann,” 527–528 (April). —— , 1802a. “Neue Versuche über den Galvanismus.” Reichs-Anzeiger (Gotha), No. 66, 813–820 (not seen; cited after Ritter 1968, 102). Reprinted in Ritter 1806a, 2, 270–284 (not seen), from which pp. 272–279 are reprinted in Ritter 1968, 84–89. —— , 1802b. “Versuche über das Sonnenlicht.” [Gilbert’s] Annalen der Physik, 12, 409–415 (December). Kleinert (1984, 295 and 298, n. 24) argued that this issue appeared no earlier than April 1803. —— , 1803a. “Versuche und Bemerkungen über den Galvanismus, als erste Fortsetzung des Aufsatzes im Magazin B. IV. S. 575–661.” [Voigt’s] Magazin für den neuesten Zustand der Naturkunde, 6, 97–129, 181–215 (August and September). Reprinted in Ritter 1806a, 3, 95–157. Dated July 1803.

ØRSTED’S PRESENTATION OF OTHERS’

337

—— , 1803b. Detailed contents of both parts of Ritter 1803a. [Voigt’s] Magazin für den neuesten Zustand der Naturkunde, 6, 175–179, as part of the “Inhalt” to that issue, 175–180 (August). Reprinted in Ritter 1806a, 3, 95–98. For authorship see note 116. —— , 1804. “[Ueber galvanisch-chemische Gegenstände].” [Gehlen’s] Neues allgemeines Journal der Chemie, 3, 692–699, in section headed “Correspondenz” (December). Dated 1 December 1804. —— , 1806a. Physisch-Chemische Abhandlungen in chronologischer Folge, 3 vols. Leipzig: C. H. Reclam. —— , 1806b. “Bermerkungen zu Herschel’s neueren Untersuchungen über das Licht;—vorgelesen in der Naturforschenden Gesellschaft zu Jena, im Frühlung 1801,” first published in Ritter 1806a, 2, 81–107. —— , 1968. Die Begründung der Elektrochemie und Entdeckung der ultravioletten Strahlen. Eine Auswahl aus den Schriften des romantischen Physikers. Ed. Armin Hermann. “Ostwalds Klassiker der exakten Wissenschaften,” N. S., 2. Frankfurt am Main: Akademische Verlagsgesellschaft. —— , 1988. Der Physiker des Romantikerkreises Johann Wilhelm Ritter in seinen Briefen an den Verleger Carl Friedrich Ernst Frommann. Ed. Klaus Richter. Weimar: Hermann Böhlaus Nachfolger. Rosenberger, Ferdinand, 1882–1890. Die Geschichte der Physik in Grundzügen mit synchronistischen Tabellen der Mathematik, der Chemie und beschreibenden Wissenschaften sowie der allgemeinen Geschichte. 3 vols. Braunschweig: Friedrich Vieweg & Sohn. Schimank, Hans, 1933. “Johann Wilhelm Ritter. Der Begründer der wissenschaftlichen Elektrochemie. Ein Lebensbild aus dem Zeitalter der Romantik.” Deutsches Museum. Abhandlungen und Berichte, Jg. 5, Hft. 6, 175–203. Also published separately, Berlin: VDI-Verlag. Schlegel, Friedrich, 1890. Friedrich Schlegels Briefe an seinen Bruder August Wilhelm. Edited by Oskar F. Walzel. Berlin: Speyer & Peters. Schuster, Johann Constantin, 1807. System der dualistischen Chemie des Prof. Jakob Joseph Winterl dargestellt von Johann Schuster M.D. 2 vols. Berlin: Frölich’sche Buchhandlung. Seebeck, Thomas Johann, 1825a. “Magnetische Polarisation der Metalle und Erze durch Temperatur-Differenz.” Abhandlungen der Königlichen Akademie der Wissenschaften zu Berlin. Aus den Jahren 1822 und 1823. Abhandlungen der physikalishen Klasse, 265–373, pl. I–II, 2 folded tables. Edited extracts of lectures read on 16 August 1821, 18 and 25 October 1821, and 11 February 1822. —— , 1825b. Ueber die Magnetische Polarisation der Metalle und Erze durch Temperatur-Differenz. Berlin: Gedruckt in der Druckerei der Königl. Akademie der Wissenschaften. [i]+109+[1] p., 2 pl., 2 folded tables. —— , 1826. “Ueber die magnetische Polarisation der Metalle und Erze durch Temperatur-Differenz. (Ausgezogen aus den so eben erschienenen Denkschriften für 1822 und 1823.).” [Poggendorff’s] Annalen der Physik und Chemie, 6 (= Annalen der Physik, 82), 1–20, 133–160, 253–286, pl. III and V (January, February, March). Snelders, H. A. M., 1970. “The Influence of the Dualistic System of Jakob Joseph Winterl (1732–1809) on the German Romantic Era.” Isis, 61, 231–240. Stauffer, Robert C., 1957. “Speculation and Experiment in the Background of Ørsted’s Discovery of Electromagnetism.” Isis, 48, 33–50. Szökefalvi-Nagy, Zoltán, 1971. “Leben und Werk von J. J. Winterl (1732–1809).” NTM. Zeitschrift für Geschichte der Naturwissenschaften, Technik und Medizin, 8, 37–45. Teuber, Jan, 2002, ed. Højdepunkter i dansk naturvidenskab. Copenhagen: Gads Forlag. Tilloch, Alexander, 1821. “Magnetism.” The Philosophical Magazine and Journal, 58, 462 (December). In the section “Intelligence and Miscellaneous Articles.” Published anonymously; editor’s authorship assumed. Winterl, Jakob Joseph, 1800. Prolusiones ad chemiam saeculi decimi noni. Budæ: Typis ac sumptibus Typographiæ Regiæ Universitatis Pestinensis. [3] + xii + 270 p. Preface dated 25 July 1800. —— , 1803. Accessiones novae ad prolusionem suam primam et secundam. Budæ: Typis ac sumptibus Typographiæ Regiæ Universitatis Pestinensis. Paged continuously with Winterl 1800, pp. 271–467. Not seen. —— , 1804. Darstellung der vier Bestandtheile der anorganischen Natur. Eine Untersuchung des ersten Theiles seiner Prolusionen und Accessionen. Trans. Johann Constantin Schuster. Pref. Johann Wilhelm Ritter. Jena: Friedrich Frommann. Not seen. —— , 1806. “Winterl’s Replik gegen eine Kritik seines Systems in d. allgem. Hallischen Literaturzeitung 1806. No. 44. S. 345–350; No. 45. S. 353–360.” [Gehlen’s] Journal für die Chemie und Physik, 1, 313– 346 (August). At the end in a footnote (345–346) is an extract from a letter from Winterl to Gehlen dated 19 July 1806. —— , 1808. “Kritik der Hypothese, welche das gegenwärtige Zeitalter der Naturwissenschaft (Physik, Chemie und Physiologie) zum Grunde legt.” [Gehlen’s] Journal für die Chemie, Physik und Mineralogie, 6, 1–35, 201–270 (January and February).

338

K. L. CANEVA

Wollaston, William Hyde, 1802. “A Method of examining refractive and dispersive Powers, by prismatic Reflexion.” Philosophical Transactions of the Royal Society of London, [92], Pt. 2, 365–380, pl. XIV. Read 24 June 1802. —— , 1803. “A Method of examining refractive and dispersive Powers, by prismatic Reflection.” [Nicholson’s] Journal of Natural Philosophy, Chemistry, and the Arts, 8vo ed. [i.e. N.S.], 4, 89–100, pl. IV (February). Abridged from Wollaston 1802. —— , 1804. “On certain Chemical Effects of Light.” [Nicholson’s] Journal of Natural Philosophy, Chemistry, and the Arts, 8vo ed. [i.e. N. S.], 8, 293–297 (August). —— , 1811. “Ueber gewisse chemische Wirkungen des Lichts.” [Gilbert’s] Annalen der Physik, 39 (= N.S., 9), 291–299 (November). Freely translated by Gilbert from Wollaston 1804, with comments. Yelin, Julius Conrad von, 1823a. Der Thermomagnetismus in einer Reihe neuer elektromagnetischer Versuche. Munich: no publisher. 12 p., 1 pl. Dated 29 April 1823; based on two lectures given at the Königliche Akademie der Wissenschaften in Munich, 12 and 26 April 1823. —— , 1823b. “Neue Versuche über die magneto-motorische Eigenschaft der bisher so genannten unmagnetischen Metalle.” [Gilbert’s] Annalen der Physik, 73, 361–364, pl. V (April). —— , 1823c. “Der Thermo-Magnetismus der Metalle, eine neue Entdeckung.” [Gilbert’s] Annalen der Physik, 73, 415–429, pl. V (April). Read at the Königliche Akademie der Wissenschaften in Munich, 12 April 1823.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY ROBERTO DE ANDRADE MARTINS

1. INTRODUCTION Magnetochemistry is the study of the effect of magnetic fields on chemical reactions. The subject received its name in the early 20th century1 but the search for such an influence began one century earlier. In the very beginning of the 19th century, after the invention of Volta’s pile and before the discovery of electromagnetism, several researchers were looking for effects of magnets on chemical reactions. One of the reasons behind this search was the evident analogy between electricity (or galvanism) and magnetism. Volta’s pile and magnets have opposite poles that exhibit attraction or repulsion. Were there any other equivalent properties? As Volta’s device could produce chemical effects, several authors expected to find similar influences of magnetism. However, as this paper will attempt to show, there were other grounds for this investigation. One of the researchers who reported chemical effects produced by magnets was Johann Wilhelm Ritter. His claims were announced by his friend Hans Christian Ørsted, who seemingly accepted his ideas and experimental results. However, other researchers could find no such effect, and Ritter’s findings were soon discredited. After the discovery of electromagnetism there arose a new wave of positive reports concerning chemical effects of magnetism, but doubts were again cast on those effects. For several decades there was a disagreement between experimental reports and it was not altogether clear whether a magnet could indeed incite any chemical change.

1

The earliest book on this subject was probably Edgar Wedekind’s Magnetochemie, published in 1911. Although the name “magnetochemistry” was not used in the early 19th century, we can use it without fear of the terrible charge of anachronism, because we find it in Lorentz Oken’s Lehrbuch der Naturphilosophie: “Magnetism and chemical action [Chemismus] are the main generating agencies for the solid nucleus of the Earth, which is build by both of them. The process of constructing the Earth is a magneto-chemical one.” Lorentz Oken, Lehrbuch der Naturphilosophie (Jena: Friedrich Frommann, 1831), p. 139. Notice that Lorentz Oken, or Ockenfuß (1779–1851) was a naturalist who embraced the philosophical school created by Schelling. He held the chair of Medicine at the University of Jena and published in 1809–1810 the first edition of his work Lehrbuch der Naturphilosophie, where he stressed the importance of polarity and the unity between galvanism and the vital force. I am grateful to Dr. Andreas Kleinert for calling my attention to this book.

339 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 339–385. © 2007 Springer.

340

R. DE A. MARTINS

This paper will present the early history of magnetochemistry, during the first three decades of the 19th century, with special emphasis on Ritter’s work and Ørsted’s involvement with this subject. This particular episode will then be discussed in the framework of the philosophical context of that time. It will be shown that there was a strong influence of Schelling’s Naturphilosophie upon Ritter’s and Ørsted’s early views on this subject and that Ritter’s search for magnetochemical effects cannot be understood without taking into account this philosophical basis.

I. EXPERIMENTAL MAGNETOCHEMISTRY 2. THE EARLIEST REPORTS ON MAGNETOCHEMISTRY Electricity and magnetism exhibit several well-known similarities. They can act at a distance, and both can produce attraction and repulsion. It was natural to think that there could be a deeper relationship between them, and towards the end of the 18th century this led the Bavarian Academy of Science to propose the following prize question (1774–1776): “Is there a true physical analogy between electric force and magnetic force?” The result of the competition was published in Van Swinden’s book, Analogie de l’éléctricité et du magnétisme, where one can find descriptions of the magnetic effects of thunderbolts side by side with curious experiments, such as G. W. Schilling’s claim that eels are attracted by magnets.2 Also recall that, around this time, Franz Anton Mesmer’s demonstrations of “magnetic” phenomena upon human beings was very influential for several years and helped to direct the attention of the researchers to the relations between life and physical forces. After the discovery of galvanism, attempts were made to find fresh correspondences between magnetism and the new phenomenon. According to Pierre Sue, Richard Fowler observed around 1796 that a magnet could induce muscular contractions, but afterwards he noticed that the same effect occurred with a non-magnetic iron bar.3 In 1797 Alexander von Humboldt (1769–1859) published his book Über die gereitzte Muskel- und Nervenfaser, where he presented and discussed several galvanic phenomena. Among them, he referred to some experiments made by Ritter, who excited contractions in frogs with magnets.4 He produced a galvanic arc with two pieces of iron and observed no twitching of the frog. He replaced one of the iron pieces by a magnet and there was an immediate twitching of the frog. He also used a chain with iron and steel and observed no effect, but when the iron or steel piece was connected to a magnet, there were strong effects. “Both experiments prove sufficiently that the magnetic steel in the galvanic chain works differently from steel or iron. This confirms Ritter’s experiments.”5 Other effects 2

3

4 5

J. H. Van Swinden, Analogie de l’éléctricité et du magnétisme ou récueil de mémoires, couronnés par l’Académie de Bavière, 3 vols. (La Haye/Paris: La Compagnie/Veuve duch*esne, 1785), vol. 1, p. 436. Pierre Sue, Histoire du galvanisme et analyse des différens ouvrages publié sur cette découverte, depuis son origine jusqu’à ce jour, 2 vols. (Paris: Bernard, an X [1802]), vol. 1, p. 207. Ritter did not publish any account of his early work on this subject. A. von Humboldt, Versuch über die gereitzte Muskel- und Nervenfaser, 2 vols. (Posen: Decker und Compagnie, 1797), vol. 2, p. 189.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

341

were however difficult to explain. When he used two similar strong magnets, no twitching occurred when the unequal magnetic poles were attached to one another, but there were contractions when the equal magnetic poles were in contact and, in this case, there was no heterogeneity that could explain the effect. In the French translation of Humboldt’s work, published two years later, he denied any direct influence of magnetism upon galvanism, but then adds: “We have certainly the right to think, according to very strong analogies, that even a weak magnet, when it is put close to a living animal or vegetable, changes the effects of its vitality and produces the acceleration of its nutrition, the general motion of fluids and other vital functions.”6 Humbolt’s views about the relation between electricity and magnetism was inconstant. He denied that the nervous and magnetic forces were of the same nature, but accepted that magnetism can influence several physiological phenomena. He admitted that Mesmer’s “magnetic” phenomena could be spurious, but that “we cannot infer from this that the application [of hands] do never produce physical effects.”7 He also wrote: It seems that the animal fibers have a property analogous to that of a magnet. In the dance of Saint Gui, the contracted muscles loosen as soon as they are touched with an iron bar. Other metals are as ineffectual as glass or wax, as reported by Scherer. This is an important discovery; but we should not conclude from this that it is the magnetic force which moves the muscles.8

At some places, he returns to the idea of a fundamental unity between galvanism, electricity, and magnetism: Perhaps the galvanic, electric and magnetic fluids have many mutual connections and only differ from one another as blood, milk and the juices of the plants, for instance. It may occur that the galvanic, electric and magnetic phenomena do not depend on particular substances, but only on the special proportions in the parts that constitute the animal body.9

In 1800, Ludwig Achim von Arnim10 published a paper on magnetism where he referred to Ritter’s experiments and tried to observe chemical effects of magnetism. Arnim reported that the two magnetic poles exhibited different oxidation phenomena. Arnim covered the two poles of a magnet with iron caps (“armatures”), and noticed that when they were moist, the North pole of the magnet and the armature at the South pole suffered a stronger oxidation. This difference in oxidation seemed to him to explain Ritter’s observation that two iron needles would produce

6

7 8 9 10

A. von Humboldt, Expériences sur le galvanisme et en général sur l’irritation des fibres musculaires et nerveuses. Traduction de l’allemand. (Paris: Didot jeune, 1799), p. 115. Humboldt, Expériences sur le galvanisme…, p. 529. Humboldt, Expériences sur le galvanisme…, pp. 453–454. Humboldt, Expériences sur le galvanisme.…, p. 454. Karl Joachim (“Achim”) Friedrich Ludwig von Arnim (1781–1831) studied natural sciences in Halle and Göttingen and medicine in Jena. He became a physician but never pursued this job. He was later to become a famous writer of the Romantic school. He is known for the volumes Des Knaben Wunderhorn (The Boy’s Magic Horn), published in 1806–1808, containing 600 folk songs he collected with his friend Clemens Brentano (1776–1842). This work strongly influenced the Grimm brothers (Jacob and Wilhelm) who began collecting folktales after reading their book. Arnim also published historical novels, such as Owen Tudor (1809) and Isabella of Egypt (1812).

342

R. DE A. MARTINS

galvanic effects. There are other curious effects described by Arnim. For instance: he reported that upon magnetisation the North pole of an artificial magnet becomes heavier, and the South pole becomes lighter.11 The relation between oxidation and galvanic effects had already been ascertained by several researchers and was discussed in Arnim’s next paper. Trial and error had shown that different metal pairs produced different galvanic effects. In the case of a silver-zinc pile the zinc pieces soon became oxidised, while silver exhibited little oxidation. It was soon suggested that the galvanic effect of a metal pair depended on their different oxidation properties. In order to obtain the strongest effect, the two metals had to exhibit the largest possible difference in their affinities for oxygen. Accordingly, several authors presented lists of metals disposed in the order of their oxidation, and the farther were the metals in the list, the stronger was supposed to be the effect of the pair. Arnim presented his own list where he emphasised that the opposite poles of a magnet exhibited different oxidation: gold—silver—mercury— copper—brass—tin—lead—iron—magnet—pyrolusite—zinc.12 According to Arnim, a magnet would suffer stronger oxidation than iron. The two poles, however, would suffer different effects. He reported that the difference in oxidation of the two poles of a magnet was clearly seen when it was put in an infusion of cress seeds. In a single night the South pole became black, while the North pole would remain bright.13 Although he did not attempt to produce a Voltaic pile using magnets, this possibility was clearly implied by his analysis and comments on the different properties of the opposite magnetic poles. Following Arnim’s work, August Friedrich Lüdicke (1748–1822) attempted to build a battery using a series of magnets.14 Lüdicke remarked that the substances of electricity and galvanism seemed the same, because both can be conducted and stored in the same bodies. The magnetic substance, on the other hand, behaves in a different way, and therefore one should hardly expect that magnetic batteries would work.15 Without any strong expectation, however, he made a trial. He used 50 pieces of magnetic iron. The “friendly” (that is, opposite) poles of successive pieces were put in contact to one another, with pieces of paper wet in salt water between them.16 The extremities of this pile were put in glass tubes, connected through a vessel full of water. The experiment began at 7 o’clock in the evening. One hour later Lüdicke observed eight very small bubbles at the North pole, and no bubbles at the South pole. Two hours later, there were 11 bubbles at the North pole and only two small bubbles at the South Pole. This showed the stronger chemical effect of the North pole.17 Notice that Arnim had observed a stronger oxidation at the South pole, instead. 11

12

13 14

15 16 17

Ludwig Achim von Arnim, “Ideen zu einer Theorie des Magneten,” Annalen der Physik 3 (1800), pp. 48–64, at p. 59. Ludwig Achim von Arnim, Bemerkungen über Volta’s Sauele, Annalen der Physik 8 (1801), pp. 163– 196, 257–284, at 279. Ibid. August Friedrich Lüdicke, Versuche mit einer magnetischen Batterie, Annalen der Physik 9 (1801), pp. 375–378. Ibid p. 375. Ibid pp. 376–377. Ibid p. 378.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

343

In 1802 William Nicholson reported in his Journal of Natural Philosophy that a correspondent (Brunn) informed him that “at Vienna18 a discovery has been made, that an artificial magnet, employed instead of a Volta’s pile, decomposes water equally well as that pile and the electrical machine; whence (as they write) the electric fluid, the galvanic fluid, and the magnetic fluid are the same.”19 Nicholson added a footnote describing that he tried the experiment but was unsuccessful. He used five bar magnets in series. Their extremities were attached to iron wires that were put in the water, and he “perceived no effect.” A few months later, Lüdicke published another paper.20 Using a larger number of magnets he obtained irregular effects. With cold water there occurred no bubbles, and with warmer water sometimes there were more bubbles at the North pole, and at other times there were less bubbles at this pole. He also remarked that using twice the number of magnets there was only a very small increase of the effect, and that it was impossible to notice any difference between the oxidation of the opposite polar surfaces. He concluded: “Thus I assume that these connected magnetic pieces may have worked here probably only as good heat conductors, and not by a kind of Galvanism.”

3. RITTER’S MAGNETOCHEMICAL EXPERIMENTS In 1803, Ritter returned to the study of magnetism. His experiments were published by his friend Ørsted, who was then in France. Ørsted recalled that “the phenomena of magnetism have frequently been compared with those of electricity, and many facts seem to justify the comparison”, but he remarked that up to that time the facts had been inconclusive.21 Mr. Ritter’s first experiments with the magnet concerned frogs. He found that a magnetic iron wire produced, with another non-magnetic wire, a galvanic palpitation in these animals. He noticed that the South pole produced stronger palpitations than the non-magnetic iron, and that the North pole excited weaker ones. Having always noticed that the metals that underwent stronger oxidation produced more powerful palpitations, he concluded that the South pole had a stronger affinity for oxygen than iron, and that the capacity of oxidation of the North pole was lower than that of iron.22

Ritter tested this conclusion by submitting a magnetised iron wire to weak nitric acid. According to Ørsted’s account, he noticed that the South pole was much 18

19

20

21

22

Maybe he was referring to Lüdicke’s experiments, but Lüdicke did not work at Vienna, but at Meissen. William Nicholson, “Scientific news,” Journal of Natural Philosophy, Chemistry, and the Arts [series 2] (1802), pp. 234–236, at 234. August Friedrich Lüdicke, “Fortsetzung der Versuche mit verbundnen Magnetstählen, und ein paar Bemerkungen zu Volta’s Sauele,” Annalen der Physik 11 (1802), pp. 114–119. Hans Christian Ørsted, “Expériences avec la pile électrique faites par M. Ritter, à Jena; communiquées par M. Ørsted,” Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts 57 (1803), pp. 401–405, at 406. Ibid. p. 406. Cf. Hans Christian Ørsted, “Experiments on magnetism; by Mr. Ritter, of Jena. Communicated by Dr. Ørsted, of Copenhagen,” Journal of Natural Philosophy, Chemistry, and the Arts, 8 (1804), pp. 184–186.

344

R. DE A. MARTINS

more strongly attacked than the other one by the acid.23 It was soon surrounded by a larger oxide deposit than the North pole.24 He made other experiments comparing the speed of oxidation of three iron wires—two of them being magnetized and the other one non-magnetic. The South pole began to exhibit oxidation before the others, and next the non-magnetised iron, and at last the North pole. It seems that it was not easy to reproduce this experiment: This experiment requires much care. The surface of the water should be covered with fresh almond oil, to avoid the admission of air. It is also necessary to avoid exposing to sunlight one of the flasks more than the others.25

The different chemical reactions could also be observed by the use of an infusion of litmus [tincture de tournesol]. The water became acid as the iron wires were oxidized, and this could be seen by observing the colour of the solution. The South pole produced a stronger red colour than the other wires, showing that it was undergoing a stronger oxidation. The effect, however, was very weak, and it was necessary to wait for more than 8 days to notice the colour change. It was advisable to add some acetic acid to the water, in such a way that the litmus infusion would be close to the point of changing from blue to red. Contact between the water and air could destroy the effect.26 Ritter also attempted to build a battery with magnets. He used 120 magnetised iron wires in series, with opposite poles close to one another but separated by a globule of water. However, the arrangement did not produce the expected effects, but Ritter did not regard this negative result as a refutation of his guiding hypothesis: “However, the clever author did not abandon his hope for composing a magnetic battery.”27 In December 1805 Ritter presented to the München Academy of Sciences a new paper where the equivalence of electricity and magnetism received an ostensibly full confirmation. He reported that he had finally succeeded in building a magnetic battery that could produce the same effects as a voltaic battery. His main results were: 1. Each magnet is equivalent to a couple of heterogeneous metals. The different poles are respectively associated to the two dissimilar metals. 2. Consequently each magnet, like these metals, produces electricity. One of the poles gives positive electricity, and the other one negative electricity. 3. A series of magnets also constitutes in analogous circ*mstances a voltaic battery, as a series of pair of different metals; and in this manner the author 23

24 25 26 27

The converse influence of chemical attack upon magnetism had been reported by Tiberius Cavallo, who claimed that after iron was attacked by acids it had a stronger effect upon a magnetic needle: Tiberius Cavallo, “Magnetical experiments and observations,” Philosophical Transactions of the Royal Society of London 76 and 77 (1785 and 1786), pp. 62–80, 6–25. Afterwards Ruhland reported a similar effect: Reinhold Ludwig Ruhland, Vermischte Bemerkungen electrischen und magnetischen Inhalts, Journal für Chemie und Physik 11 (1814), pp. 16–25. Ørsted, Expériences avec la pile électrique…, pp. 406–407. Ibid p. 407. Ibid p. 407. Ibid pp. 408–409.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

345

demonstrated, by means of the electrometer, the electricity produced by the poles of this series of magnets. 4. With this power the battery of magnets exercises upon living or inanimate bodies, the same effects as a voltaic column of equal strength.28 These experiments demonstrate that, in magnetised iron, the South pole yields positive electricity, and the North pole the negative. On the contrary, in the magnetised steel, the north pole yields positive electricity, and the south pole yields negative. The same inverse distribution is observed in the influence of magnetic polarity upon oxidation of the magnetised body. In iron, the South pole undergoes stronger oxidation, and the North pole a weaker one. In magnetised steel, on the contrary, the North pole undergoes stronger oxidation, and the South pole a weaker effect.29

The Italian translator of the letter reporting Ritter’s results (who was probably Carlo Amoretti, the editor of the journal), added a note commenting that there was also an electric and magnetic polarity in fruits and seeds. The part of the seed where the roots are to appear has positive electricity, and the opposite part is negative. When a pine cone was suspended by a silk thread, inside a glass container, it would turn when a magnet was approached. The South pole of the pine cone corresponded to the part where the roots would appear upon germination. However, when the pine cone was stripped of its hard sheath, this side behaved as the North pole. It is likely that many researchers unsuccessfully attempted to repeat several of Ritter’s experiments, but no public criticism had come to light. In 1807, however, Ritter’s work met severe disapproval. Paul Erman (1764–1851) published two lengthy papers, where he presented a detailed experimental criticism of Ritter’s work. There were several points under attack: Ritter’s work on atmospheric electricity and the aurora borealis; his claims concerning the electrical poles of the Earth; the attraction between a silver-zinc needle and a magnet; and the influence of magnetism upon chemical reactions. Contrary to the previous reports of other researchers, Erman found no difference in oxidation between the North and South poles of several magnets.30 None of his replications of the earlier experiments was successful, and he concluded that Ritter’s claims were groundless. Erman was a respected physicist, and his papers represented a serious challenge to Ritter’s claims. Erman’s work was not disputed. It helped to bury magnetochemistry for several years, together with other results reported by Ritter.

28

29 30

[Anonymous]. Extrait d’une lettre écrite de Munich au Prof. Pictet sur quelques expériences galvanico-magnétiques, faites récémment par Mr. Ritter, Bibliothèque Britannique, ou Recueil. Sciences et Arts 31 (1806), pp. 97–100. This notice was reproduced in a number of journals: [Anonymous]. Aus dem Intelligenzblatte der Allgem. Litterat. Zeit. Den 5ten Febr. 1806, Annalen der Physik 22 (1806), pp. 223–224; “Extract of a letter to professor Pictet, from a Correspondent at Munich, upon some galvanico-magnetic experiments recently made by M. Ritter,” Philosophical Magazine 25 (1806), pp. 368–369; Estrato d’una lettera scritta da Monaco in Baviera al Sig. Prof. Picted di Ginevra su alcuni sperimenti galvanico-magentici fatti recentemente dal Sig. Prof. Ritter, Nuova Scelta d’Opuscoli 1 (1806), pp. 334–336. Ibid. Paul Erman, Beitraege über electrisch-geographische Polaritaet, permanente electrische Ladung, und magnetisch-chemische Wirkungen, Annalen der Physik 26 (1807), pp. 1–35, 121–145, at 141–142.

346

R. DE A. MARTINS 4. MAGNETOCHEMISTRY AFTER THE DISCOVERY OF ELECTROMAGNETISM

In 1820 Hans Christian Ørsted published his discovery of the magnetic effect of galvanic currents, and this finding produced abundant progeny.31 Some of the works that followed Ørsted’s—such as Ampère’s researches on electrodynamics and Seebeck’s discovery of thermoelectricity—are well known. Some others have not been incorporated into mainstream science and have been forgotten. Soon after the announcement of Ørsted’s discovery, Dominique François Arago reported that an iron wire wound around a cylinder and connected to a galvanic apparatus produced strong magnetic effects. It immediately occurred to Augustin Fresnel that an inverse effect could also exist: perhaps a magnet could produce a voltaic current in a metallic wire coiled around the magnet. His first trials seemed to manifest positive results: the end of the wire that he expected to become positive underwent strong oxidation in water, while the other end suffered no oxidation for several days. He was therefore persuaded that magnetism had produced a voltaic current and a chemical effect. Hence, on the 6th November he presented the confirmation of his conjectures to the French Academy of Sciences.32 After a few weeks, as further experiments did not confirm his earlier findings, Fresnel concluded that the effect did not exist.33 One week after Fresnel presented his first results, C. J. Lehot claimed that he had already discovered the same effect six years before.34 He reported that an iron wire connected to the South pole of a magnet suffered much stronger oxidation in water than another similar wire attached to the North pole. The different chemical effects of the North and South poles could also be observed by the colour of a litmus infusion, which became red around the wire connected to the South magnetic pole. Lehot recalled that those experiments had already been made 20 years earlier by Ritter, and that they were cited in several works on galvanism. Soon afterwards, John Murray presented to the Royal Society of Edinburgh a paper where he described some chemical effects of magnetism. Among other phenomena, he described that magnetism would produce the reduction of silver and

31

32

33

34

Hans Christian Ørsted, Experimenta circa effectum conflictus electrici in acum magneticum (Hafniae: Schultz, 1820). Augustin Fresnel, Note sur des essais ayant pour but de décomposer l’eau avec un aimant, Annales de Chimie et de Physique [series 3] 15 (1820), pp. 219–222. Ten years later, after the discovery of electromagnetic induction, Ampère suggested that Fresnel’s experiment could have exhibited the effect of induced currents. Of course, the motion of magnets would produce only short-lived currents, but he thought that the continuous temperature changes of the magnets that must have occurred during those long-term experiments could produce significant currents. Antoine César Becquerel and Edmond Becquerel, Traité de l’électricité et du magnétisme et des applications de ces sciences a la chimie, a la physiologie et aux arts, 3 vols., (Paris: Firmin Didot, 1855–1856), vol. 1, p. 384. Yelin repeated Fresnel’s experiments and could not observe any positive effect, either: Ritter von Yelin, Ueber den Zusammenhang der Electricitaet und des Magnetismus, Annalen der Physik 66 (1820), pp. 395–411. C. J. Lehot, [M. Lehot adresse une lettre relative à l’expérience dont M. Fresnel a rendu compte à l’Académie dans la dernière séance]. Annales de Chimie et de Physique [series 3] 15 (1820), pp. 406–408.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

347

its precipitation in the form of small crystals. A non-magnetic steel wire, put in a solution of silver nitrate, produced no chemical change, but when it was attached to the north and south poles of two magnets, it soon became covered with crystals of silver. When a magnet was put in the solution of silver nitrate, “the North pole became instantly studded with brilliant pallets of silver, and formed more rapidly and more copiously round it than round the south pole.”35 New reports continued to appear. In January 1821, Ørsted’s friend Christopher Hansteen wrote a letter to Ludwig Wilhelm Gilbert (the editor of the Annalen der Physik) describing experiments that had been made a few years earlier by Hans Henrik Maschmann and himself concerning the chemical effects of magnetism.36 Maschmann, a chemistry professor at the university of Christiania, in Norway, observed in 1817 that the crystallisation of silver (forming Diana’s silver tree) from a solution of silver nitrate under the influence of metallic mercury was stronger to the North side of the glass tube he used, and conjectured that the effect could be due to the magnetic field of the Earth. Several later experiments, using both the magnetic field of the Earth and the influence of nearby magnets confirmed that the formation of Diana’s tree was faster under the influence of the North magnetic pole. He interpreted the chemical effect as due to galvanism, and concluded that galvanism and magnetism were identical. He also conjectured that magnetism could have some effect in geological phenomena.37 Maschmann communicated his discovery to his colleagues Hansteen and F. Keiser—who confirmed his findings—and also to Ørsted.38 Notice that this happened three years before the discovery of electromagnetism. Hansteen, on the other hand, declared that he wrote a paper on those experiments and sent it to Ørsted in 1819, but the paper was not returned.39 A few years later Maschmann’s and Hansteen’s papers were translated into French, when the Abbot Louis Rendu reported other chemical effects of magnetism. Together with those articles there appeared a paper by Ørsted where he described Ritter’s experiments on the chemical effects of magnets.40 Any well informed scientist, at that time, would become aware of the existence of several phenomena exhibiting a relation between magnetism and chemistry, and there were many clues pointing out that Ørsted and Ritter had something to do with that subject.

35

36

37

38 39 40

John Murray, “On the decomposition of metallic salts by the magnet,” London, Edinburgh, and Dublin Philosophical Magazine 58 (1821), pp. 380–382, at 381. Christopher Hansteen, Wiederholung und bestätigung der Versuche durch Hrn. Prof. Hansteen, Annalen der Physik 70 (1822), pp. 239–242. Hans Henrik Maschmann, “Einwirkung des Erdmagnetismus auf Auscheidung des Silbers,” Annalen der Physik 70 (1822), pp. 234–239. Ibid p. 238. Hansteen, Wiederholung unt bestätigung der Versuche…, p. 241. Louis Rendu, Influence du magnétisme sur les actions chimiques, Annales de Chimie et de Physique [series 3] 38 (1828), pp. 196–197; Hans Christian Ørsted, Experiénces de Ritter, analysées par M. Ørsted, Annales de Chimie et de Physique [series 3] 38 (1828), pp. 197–200. This paper was a partial reproduction of an article published by Ørsted 25 years earlier, but the journal did not inform the readers about that.

348

R. DE A. MARTINS 5. DIANA’S SILVER TREE

As described above, one of the claims published in the following years was the influence of magnetism upon Diana’s silver tree. The reduction of metallic salts in aqueous solution produce in special circ*mstances metallic crystals that build up a treelike (dendritic) structure. This kind of phenomenon had already called the attention of alchemists, who described the so-called “Diana’s tree” (Arbor Dianæ), built up of silver crystals: The Reign of the Moon lasts just three weeks; but before its close, the substance exhibits a great variety of forms; it will become liquid, and again coagulate a hundred times a day; sometimes it will present the appearance of fishes’ eyes, and then again of tiny silver trees, with twigs and leaves. Whenever you look at it you will have cause for astonishment, particularly when you see it all divided into beautiful but very minute grains of silver, like the rays of the Sun. This is the White Tincture, glorious to behold, but nothing in respect of what it may become.41

In the early 19th century it was found that electricity may quicken the precipitation of those “metallic trees.” The “tree of Saturn” can be produced when a copper wire attached to a zinc plate is put inside a diluted solution of neutral lead acetate. Lead precipitates in the form of small bright plaques attached to the wire, and new crystals form upon the first ones, building a treelike framework that gradually grows in the containing vessel.42 This effect was discovered by William Cruickshank, who conjectured that hydrogen produced by the Voltaic decomposition of water could reduce metals.43 He described that in the case of silver nitrate the metal precipitated in the form of small needle-like crystals building up Diana’s tree. In the same year, Richard Kirwan conjectured that crystallisation could be due to magnetic force.44 Independently, Ritter had also noticed that Volta’s pile could produce metallic “vegetations” similar to Diana’s silver tree. The negative galvanic lead was able to reduce several metals from their salt solutions to metallic crystals, and sometimes the metal crystals gather as the branches of a tree.45 As described above, Maschmann, Hansteen, and Murray had claimed that a magnet could have an effect on the formation of Diana’s tree. However, shortly after the publication of Murray’s paper it was criticised by an anonymous

41

42 43

44

45

This book of Eirenaeus Philalethes was first published as Introitus apertus ad occlusum regis palatium (Amsterdam, 1667) and translated as Secrets reveal’d: or, an open entrance to the shut-palace of the king (London, 1669). The complete electronic text of this book can be found on the Internet, in two different versions: at http://clairvision.org/EsotericKnowledge/Alchemy/Hermetic_Museum/ Open_Entrance.html and also at http://www.levity.com/alchemy/openentr.html. The citation was taken from the first electronic version. Becquerel and Becquerel, Traité d’électricité et du magnétisme, vol. 2.…, pp. 196–197. William Cruickshank, “Some experiments and observations on galvanic electricity,” A Journal of Natural Philosophy, Chemistry, and the Arts [series 1] 4 (1801), pp. 187–191. This paper was translated into German as Versuche und Beobachtungen ueber einige chemische Wirkungen der galvanischen Electricitaet, Annalen der Physik 6 (1800), pp. 360–368. Richard Kirwan, “Thoughts on magnetism,” A Journal of Natural Philosophy, Chemistry, and the Arts 4 (1801), pp. 90–94, 133–135. Hans Christian Ørsted, Ueber der neuesten Fortschritte der Physik, Europa. Eine Zeitschrift 1 (1803); reproduced in Kirstine Meyer (ed.), H.C. Ørsted, Scientific Papers (Copenhagen: Andr. Fred. Høst & Søn, 1920), vol. 1, pp. 112–131, at 117.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

349

author.46 “B. M.” repeated all experiments described by Murray. He reported that he observed no sensible difference between the influences of magnetic and nonmagnetic steel in the precipitation of silver. Murray replied and strongly protested against the anonymous attack: “If truth be the object of this writer, why does he blush to own [his name]? Is science to be a masquerade, and its friends appear in false or fictitious characters? An honest man ought to be ashamed of such a contemptible subterfuge […].”47 He conjectured that the steel used by “B. M.” could be slightly magnetic, because the only test that “B. M.” had applied was to check whether it attracted iron filings, and that test was not very sensitive. He also claimed that a magnet could precipitate silver from a solution of silver acetate, and that iron could never produce such an effect. “B. M.” answered to Murray’s reply, but did not comment on the two relevant points of Murray’s reply.48 Murray’s paper produced some polemical papers in Italy, too. Ridolfi reported that he could not repeat Murray’s results, and recalled that two other physicists (Catullo and Fusinieri) had also disconfirmed those experiments. However, two other researchers, Nobili and Merosi, claimed that they had successfully repeated Murray’s experiments.49 Maschmann’s and Hansteen’s experiments were successfully repeated by Johann Schweigger, who was studying a new kind of metallic “vegetation” produced by the reduction of copper solutions: the “Venus tree” or Arbor Veneris.50 He observed that the metallic tree grew larger towards the North. According to him, Döbereiner also obtained positive results similar to those reported by Maschmann and Hansteen.51 In the same year, Karl Kastner also reported that the reduction of metallic salts was stronger towards the North. Friedrich Dulk, however, was unable to observe any influence of magnetism on the growth of Diana’s silver tree.52

6. OTHER POSIVE RESULTS AND CRITICISM It seems that Ørsted’s demonstration that an electric current produces a magnetic effect led many authors to believe that all electric and magnetic phenomena were equivalent. The Abbot Louis Rendu published a paper where he claimed 46

47 48 49 50

51 52

B. M. “Observations on Mr. Murray’s paper on the decomposition of metallic salts by the magnet,” The Annals of Philosophy [ser. 2] 3 (1822), pp. 39–41. John Murray, “Reply to B. M.,” The Annals of Philosophy [ser.2] 3 (1822), pp. 121–123, at 121. B. M., “An answer to Mr. Murray’s ‘Reply’,” The Annals of Philosophy [ser. 2] 3 (1822), pp. 384–385. C. Ridolfi, Lettera all’editore dell’Antologia, Antologia 7 (1822), 498–501. Johann Salomo Christoph Schweigger, Ueber Cohäsion, in Abhaengigkeit von krystall-elektrischer Anziehung, Journal fuer Chemie und Physik 44 (1825), pp. 79–86, at 81. Notice that Schweigger is also classified as a Romantic physicist: see Walter Kaiser, “Symmetries in Romantic physics,” in Manuel G. Doncel, Armin Hermann, Louis Michel, & Abraham Pais (eds.), Symmetries in Physics (1600–1980) (Barcelona: Universitat Autònoma de Barcelona, 1987), pp. 77–92. Ibid p. 85. Karl Wilhelm Gottlob Kastner, Zur Geschichte des Galvanismus, Archiv fuer die gesammte Naturlehre 6 (1825), pp. 442–452, at 450; Friedrich Philipp Dulk, Ueber die chemische Einwirkung des Magnetismus, Archiv fuer die gesammte Naturlehre 6 (1825), pp. 457–467.

350

R. DE A. MARTINS

that crystallisation was an electrical phenomenon and called the attention to the similarity between needle-like metallic crystals produced in electrolysis and the arrangement of iron filings submitted to a magnet.53 Guided by this analogy, Rendu attempted to produce chemical effects attaching iron wires to the poles of a magnet.54 He used a V-shaped glass tube filled with a blue tincture of red cabbage, and introduced the iron wires in each of the branches of the tube. In about 15 minutes the liquid had turned green. It was known that acids would turn this tincture red, and alkalis would turn it green.55 Rendu communicated his result to Biot, who conjectured that the effect might be due to a chemical reaction of the iron, instead of a magnetic effect. He suggested to Rendu a new experiment that excluded chemical reaction between iron and water. The iron wires were enclosed in thin glass tubes, closed at its ends, and therefore did not touch the liquid. In the modified experiment the tincture did again become green, as in the former case, but only after 2 hours. Rendu remarked that the tincture turned red, not green, when left to itself.56 Rendu’s experiment, communicated to the Paris Academy of Sciences by Biot, called again the attention of researchers to the relation between magnetism and chemical reactions. Karl Kastner reported that he also observed an effect of magnetism on vegetable tinctures and on the crystallisation of metals in saline solutions.57 In Italy, the priest Francesco Zantedeschi repeated Ritter’s experiments and reported that a steel needle attached to the North pole of a strong magnet underwent faster attack in acidulated water than another steel needle attached to the South pole. According to this author, the effect depended on the position of the magnet: it was stronger when the North pole pointed to the North or to the West.58 This author claimed that in some of his experiments the magnet became weaker after producing chemical effects.59 He also reported that a copper wire attached to the opposite poles of a magnet and connected to a multiplier (an early type of galvanometer) exhibited an effect corresponding to an electric current.60 Gustav Wetzlar, however, could observe no influence of magnetism upon the reduction of copper sulphate by iron.61 53

54

55 56 57

58

59 60

61

Louis Rendu, “Observations qui tendent a prouver que la cristallisation de tous le corps est un phénomène électrique,” Bibliothèque Universelle des Sciences, Belles Lettres et Arts. Sciences et Arts 38 & 39 (1828), pp. 304–17, 58–72, at 310–311. Louis Rendu, Influence du magnétisme sur les actions chimiques, Annales de Chimie et de Physique [series 3] 38 (1828), pp. 196–197. Rendu, Observations qui tendent a prouver…, p. 314 Rendu, Influence du magnétisme…, p. 197. Karl Wilhelm Gottlob Kastner, Chemische Gegenwirkung des magnetischen Eisens. Nachtrag zum Vorhergehenden, Archiv fuer gesammte Naturlehre 15 (1828), pp. 336–344. Francesco Zantesechi, Nota sopra l’azione della calamita e di alcuni fenomini chimici, Biblioteca Italiana o sia Giornale di Letteratura, Scienze ed Arti 53 (1829), pp. 398–402, at 400. Ibid p. 401. Ibid p. 402. A few years later Zantedeschi was to claim that he had discovered electromagnetic induction before Faraday: Francesco Zantedeschi, Relazione delle principali scoperte magneto elettriche (Verona: Antonelli, 1834) (Opuscoli Fisici di Varii Autori, vol. 1, n. 36). Gustav Wetzlar, Ueber den elektrodynamischen Zustand, welchen Eisen und Stahl durch Beruehrung mit saurer salpetersaurer Silbersloesung oder reiner Ammoniakfluessigkeit erlangen, Journal fuer Chemie und Physik 56 (1829), pp. 206–227.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

351

After several authors had reported positive findings, the Leipzig physicist Otto Linné Erdmann attempted to ascertain whether those chemical effects of magnetism did really exist. He used very strong magnets and repeated every kind of experiment that had been previously described. He noticed that several influences could affect the observed phenomena, and stressed that it was necessary to repeat many times each experiment, in different circ*mstances.62 He noticed, for instance, that the same iron wire, cut into several pieces, exhibited points where oxidation was stronger or weaker, although they seemed exactly alike in all respects. Contact of the wires with the experimenter’s hands or with different substances also affected their attack by water and mild acids. Erdmann tested several reported effects: 1. the influence of terrestrial magnetism on the oxidation of non-magnetic iron wires; 2. the differential oxidation of the poles of magnets and magnetic iron; 3. the influence of the terrestrial magnetic field on the building of Diana’s and Saturn’s trees; 4. the influence of magnets on the same phenomena; 5. the change of colour of vegetable tinctures by magnetic action. In a large series of experiments, taking care to avoid spurious influences, Erdmann could observe no positive effect of magnetism in any of those chemical reactions. He concluded that former researchers who had reported positive effects had been mistaken. Abstracts of Erdmann’s paper soon appeared in French and in English.63 As had happened in the case of Paul Ermann’s 1807 paper, his experiments seemed convincing and were cited by several authors as a definitive proof that magnetism had no influence on chemical phenomena.

7. AFTERMATH In 1831 Jacob Berzelius described Erdmann’s researches and remarked that he had also looked for chemical effects of magnetism many years before (in 1812), but obtained only negative results.64 In 1834, in his treatise on electricity and magnetism, Antoine César Becquerel supplied a short review of this subject. He regarded Erdmann’s researches as conclusive 62

63

64

Otto Linné Erdmann, Versuche ueber den angeblichen Einfluss des Magnetismus auf chemische Wirkungen und auf den Krystallisationsprocess, Journal fuer Chemie und Physik 56 (1829), pp. 24–53, at 34. Otto Linné Erdmann, “Expériences sur l’influence presumée du magnétisme sur les effets chimiques et la marche de la cristallisation”, Bibliothèque Universelle des Sciences, des Belles-Lettres et Arts. Sciences et Arts 42 (1829), pp. 96–103; Erdmann, “On the supposed influence of magnetism in the phenomena of chemical combinations and crystallizations,” The American Journal of Science and Arts 18 (1830), pp. 395–397. Jons Jacob Berzelius, Jahresbericht ueber die Fortschritte der physischen Wissenschaften, 30 vols., (Tuebingen: H. Laupp, 1822–1851), at vol. 10, pp. 42–43.

352

R. DE A. MARTINS

and denied the existence of chemical effects of magnetism.65 Moreover, the 8th edition of the Encyclopaedia Britannica, after describing several experiments made by Ritter, Fresnel, and Maschmann, presented this final comment “Mr. Erdmann, after a very elaborate inquiry into the effects of magnets as chemical agents, came to the conclusion that the observed phenomena were due to the influence of other causes, which had not been sufficiently guarded against.”66 Not every author concluded that Erdmann’s researches were conclusive. In 1843 Leopold Gmelin, in his famous Handbuch der Chemie, presented a roll of authors who defended the existence and another list of those who denied the phenomenon, but did not state his own opinion.67 Again and again there appeared in the scientific journals several claims concerning magnetochemical effects, and an equivalent number of denials of those claims. Towards the end of the 19th century Gustav Wiedemann devoted just a few paragraphs of his treatise on electricity to the description of old works and denied the phenomenon. In the same way, Wilhelm Ostwald dismissed the old claims and classified Ritter’s work as “galvanic fantasies.”68 Most physicists and chemists had forgotten this subject towards the end of the 19th century. In the decades of 1880 and 1890, however, the study of this subject received a new impetus from both the experimental and the theoretical points of view. Indeed, in the two last decades of the 19th century some magnetochemical phenomena became well-behaved and were accepted by the scientific community. In 1881, Ira Remsen found out that a magnetic field might weaken the chemical reaction between an iron plate and a solution of copper sulfate. A few years later Paul Janet and Pierre Duhem discussed the thermodynamic interpretation of the phenomenon. As the result of theoretical analysis, it was established that there should be an electromotive force between two equal iron electrodes, if one of them is magnetised and the other is not. Therefore, in a sense, it should be possible to produce electrolysis using a magnet. Afterwards there were several attempts to detect this effect. However, the predicted electromotive force was small and experiments produced conflicting results. The first researcher who obtained regular effects, compatible with thermodynamic predictions, was the Romanian physicist Dragomir Hurmuzescu.69 In a long series of works, published from 1894 onward, he developed a successful experimental method that was reproduced by other authors, such as René Paillot. Hurmuzescu’s work was regarded as so momentous that he was invited to report his researches at the 1900 Congrès International de Physique, in Paris. After Hurmuzescu’s work, most authors agreed that the effect existed. However, 65

66

67

68

69

Antoine César Becquerel, Traité expérimental de l’électricité et du magnétisme et de leur rapports avec les phénomènes naturels, 7 vols., (Paris: Firmin Didot, 1834–1840), at vol. 1, pp. 380–386. [Anonymous], “On the influence of magnetism on chemical action,” Encyclopedia Britannica, 8th ed. (Edinburgh: Adam & Charles Black, 1857), vol. 14, pp. 41–42. Leopold Gmelin, Handbook of chemistry, translated by Henry Watts, 18 vols. (London: Cavendish Society, 1848–1872), vol. 1, p. 514. Gustav Wiedemann, Die Lehre von der Elektricität (Braunschweig: Friedrich Vieweg und Sohn, 1882–1883), vol. 3, §1125, pp. 967–968; Wilhelm Ostwald, Elektrochemie. Ihre Geschichte und Lehre (Leipzig: Von Veit, 1896), pp. 216–217. Roberto de Andrade Martins, “The rise of magnetochemistry, from Ritter to Hurmuzescu” [forthcoming].

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

353

as the effect was weak and difficult to detect, its practical importance was negligible. The subject was gradually forgotten by chemists and physicist,70 and never attracted the attention of historians.71

8. ØRSTED’S OPINION In 1830, in the article on thermoelectricity he wrote for the Edinburgh Encyclopaedia, Ørsted described approvingly Maschmann’s and Hansteen’s findings and acknowledged that they could be regarded as forerunners of the discovery of electromagnetism, because of their magnetochemical experiments: Two or three years before the discovery of electromagnetism, Professor Maschmann at Chrisiania, in Norway, observed that the silver tree, formed in a solution of nitrate of silver, when put in contact with mercury, (the arbour Dianæ) takes a direction towards the north; and the celebrated Professor Hansteen found that this direction can likewise be determined by a great magnet. As the metallic precipitation is also of galvanical nature, this observation may be considered as one of the precursors of electromagnetism.72

In the same article, on the other hand, Ørsted denied Ritter’s early results: Joh. Will. Ritter, already mentioned, pursued a great number of researches upon the analogy of magnetism and electricity. He had in the year 1801 made a series of very delicate experiments upon the galvanical difference between the two magnetical poles of a steel needle. The result deduced from his experiments was, that the southern extremity of the needle was more oxidable than the northern, and that the galvanical effect of two magnetical needles upon a frog was such, that the south pole acted as the more oxidable, the north pole as the less oxidable metal. It is now acknowledged, that he has been led into error by the difference which a small disparity in the polish of the metal can produce, and which he employed insufficient means to avoid. […] The precipitation with which Ritter published these and some other erroneous statements, has thrown a shade over the name of this unhappy but ingenious philosopher, who has enriched science with several discoveries of great importance, and whose profound yet obscure ideas in many cases have anticipated the discoveries of future times.73

It is difficult to ascertain when Ørsted came to reject Ritter’s results. In 1812, in his Ansicht der chemischen Naturgesetze, Ørsted still accepted that the magnetic South pole suffers a stronger oxidation.74 He was aware that Ritter’s experiments had been criticised, but he accepted their main result: the establishment of a relation between electricity and magnetism. 70

71

72

73 74

One can still find a description of this subject in Bhatnagar and Mathur’s textbook, Physical principles and applications of magnetochemistry (1935), but Selwood’s Magnetochemistry (1943) does not describe this phenomenon. Shanti Swarupa Bhatnagar and K. N. Mathur, Physical principles and applications of magnetochemistry (London: MacMillan, 1935); Pierce Wilson Selwood, Magnetochemistry (New York: Interscience, 1943). In the subject volumes of the Isis Cumulative Bibliography it is possible to find an empty entry for “magnetochemistry.” Hans Christian Ørsted, “Thermo-electricity,” in David Brewster (ed.), The Edinburgh Encyclopaedia, 18 vols. (Edinburgh, 1830), vol. 18, pp. 573–589, at 575. Ibid p. 574. Ørsted, Ansicht der chemischen Naturgesetze durch die neueren Entdeckungen gewonnen, in Hans Christian Ørsted, Scientific papers, edited by Kirstine Meyer (Copenhagen: Andr. Fred. Høst & Søn, 1920), vol. 2, pp. 35–169, at 148.

354

R. DE A. MARTINS The remaining similarities between magnetism and electricity are so great that we need only remove the apparent contradictions in order to accept the identity of the forces in them. […] Ritter has also found that magnetized iron wire is less oxidizable at its northern end and more oxidizable at its southern end than iron, but iron or soft steel must be used here because harder steel produces less activity and, in fact, in the reversed order due to its poorer conduction and its corresponding smaller quantity of force. Under similar conditions, muscular contractions are also induced in a prepared frog if two opposite poles of a magnetized iron wire are connected to it in such a way that a closed circuit can be formed. The wires must be magnetized by means of relatively strong magnets. These experiments are still somewhat disputed by physicists, but so many have been successful that it is not easy to assume a false conclusion. […] Therefore, all the functions which can be demonstrated in electricity can also be observed in magnetism: attractions and repulsions, chemical difference, effects on the living animal body, the production of light.75

It is likely that Ritter’s magnetochemical experiments influenced Ørsted’s acceptance of the “identity” between electricity and magnetism, and hence had some bearing on his discovery of electromagnetism.

II. THE PHILOSOPHICAL BACKGROUND After this brief historical description, one might be tempted to interpret this confusing episode as the result of mere empirical exploration, misguided by vague analogies between electricity and magnetism. The second part of this paper will attempt to reveal that a deep commitment to Naturphilosophie guided Ritter’s researches on magnetochemistry. At the time when he developed his magnetochemical studies, Ritter was carrying out a search for relations between the polarities of different types of “forces.” This search was grounded upon a belief in the deep unity of all natural forces, and was strongly influenced by Schelling’s Naturphilosophie. It will be also shown that Ørsted had similar beliefs, at that time.

9. RITTER ON THE PHYSIOLOGICAL EFFECTS OF GALVANISM Around 1800, Ritter was studying several different phenomena, paying special attention to their polarity—that is, the existence of opposite extremes in each of them. Volta’s pile provided a new impetus to his research, and he began to compare the positive/negative polarity of the galvanic poles to other oppositions—both in phenomena produced by galvanism and in other fields. Armin Herman has provided an illuminating description of Ritter’s discovery of ultraviolet light.76 He emphasised that Ritter’s guiding principle was the search for complementary or opposite aspects in light. He had heard about 75

76

Ørsted, “View of the chemical laws of nature obtained through recent discoveries,” (1812), in Selected scientific works…, p. 379. Armin Hermann, “Unity and metamorphosis of forces (1800–1850): Schelling, Ørsted, and Faraday,” in Manuel Doncel, Armin Hermann, Louis Michel, and Abraham Pais (eds.), Symmetries in Physics (1600–1980) (Barcelona: Universitat Autònoma de Barcelona, 1987), pp. 51–62, at 58. Regarding ultraviolet light, I will use anachronistic terms here, for the sake of conciseness.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

355

William Herschel’s discovery of the infrared and knew that this radiation could be detected with a thermometer, because it produced heat. He expected to find another invisible radiation, at the other end of the spectrum, that would produce cold, but no such effect was observed. He was then led to search for another property of the hypothetical radiation, and he was successful to observe that it could produce the chemical reduction of silver chloride. His final interpretation was that the fundamental property of the infrared rays was not the production of heat, but an oxidation effect, opposite to the reduction effect produced by ultraviolet rays.77 In a paper published in 1803 presenting “A review of the latest advances in physics”, Ørsted strongly emphasised Ritter’s recent findings and depicted the establishment of the polarity of light as the most important aspect of the event we call “the discovery of ultraviolet radiation”: Violet light is the most deoxidizing among the light rays, which Scheele’s experiments have taught us. Herschel demonstrated to us that red light is accompanied by the greatest warming, and at the same time he proved that next to the red light there are invisible rays which possess an even greater warming ability. However, these discoveries were still quite isolated, without any connection to the remaining phenomena until Ritter discovered that there are invisible rays on both sides of the spectrum; that those on the violet side cause deoxidation, those on the red oxidation, and that the rays promote oxidation more, the closer they are to the red; similarly, they promote deoxidation more, the closer they are to the violet.78

Let us consider another example: Ritter’s researches of the effect of the voltaic poles upon the sense organs. During his early studies of galvanic phenomena, in 1792, Volta had reported that a metallic couple applied to the tip and the middle of the tongue produced an acid taste, and that the same metallic couple applied to the eye would produce visual effects.79 John Robinson also reported, in 1793, that a couple of zinc and silver, applied to the eye, produced a luminous flash. Stimulated by Robinson’s results, Richard Fowler studied the effects produced by metallic pairs upon the several sensorial organs. He noticed that they could produce strong effects upon the ear.80 No smell was produced, however, when the galvanic arc was applied to the nose. Humboldt and other researchers confirmed this result.81 77

78 79

80

81

Besides noticing that the invisible radiation close to the violet end of the spectrum produced the reduction of silver nitrate, Ritter observed that a sample with slightly reduced silver nitrate would recover its white colour when it was submitted to red light, or to infrared radiation. Ørsted, Expériences avec la pile électrique…, pp. 409–410. So, he concluded that the solar spectrum is followed by invisible rays that produces oxygenation on the red side, and reduction on the violet side. Ørsted, “A review of the latest advances in physics,” (1803), in Selected scientific works…, p. 107. Marcello Pera, The Ambiguous Frog: the Galvani-Volta controversy on animal electricity, translated by J. Mandelbaum (Princeton, NJ: Princeton University Press, 1992), pp. 107, 109. Pera observes (p. 182) that this phenomenon had already been described in 1767—many years before the discovery of Galvanism—by Johann Georg Sulzer. L. S. Jacyna, “Galvanic influences: themes in the early history of British animal electricity,” in Marco Bresadola and Giuliano Pancaldi (eds.), Luigi Galvani International Workshop. Proceedings (Bologna: Universita de Bologna, 1999), pp. 167–185, at 167, 171–172. Paul Fleury Mottelay, Bibliographical History of Electricity and Magnetism (London: Charles Griffin, 1922), pp. 307, 311, 333.

356

R. DE A. MARTINS

The invention of Alessandro Volta’s new apparatus (the pile) was communicated in a letter to Joseph Banks. There he described several physiological effects produced with his instrument. These included shocks, the pain produced by the voltaic current upon wounds, the acid or bitter taste produced on the tongue, and luminous sensations on the eye. When he applied to his ear the wires connected to the pile, he felt a very strong shock and some indefinite noise.82 No specific sensation was produced on the nose. The effects were stronger but altogether similar to those produced by the galvanic arc. Most researchers, after confirming Volta’s experiments, did not pursue those physiological investigations. Ritter, however, was not content with those results. He applied metallic wires attached to the voltaic pile to his eyes, skin, nose, etc. and described the observed effects. His results were published in 1801. They were much more definite than those of the former authors. Instead of describing just a flash of light when he applied the pile to his eye, his account was full of details. When the zinc pole (positive lead) of the battery was linked to his eye, Ritter observed first a flash and then a blue colour. If the contact was kept for some time, he observed that the objects he looked at seemed smaller and less distinct. When contact was broken he saw again a flash and then a red colour. Opposite effects were described when he connected his eye to the silver or copper pole of the pile (negative lead). After the initial flash he saw a red colour, and the objects he looked at seemed larger and more distinct. After contact was broken, he saw a blue colour.83 In the case of the other sense organs Ritter also arrived at new results. The following table summarises Ritter’s conclusions about the physiological effects of the two different kinds of voltaic electricity:

Positive pole

Negative pole

1. 2. 3. 4.

1. 2. 3. 4.

expansion of the tissues; sensation of heat; stronger pulsation; the eyes see a red colour, larger and more distinct images; 5. the tongue perceives an acid taste; 6. the nose has a reduced sense of smell, as that produced by acids; 7. the ears sense a grave sound.

contraction of the tissues; sensation of cold; weaker pulsation; the eyes see a blue (or violet) colour, smaller and less distinct images; 5. the tongue perceives an alkaline taste; 6. the nose gets the impression of an ammoniacal smell; 7. the ears sense an acute sound.

Of all those distinctions, only the acid and bitter tastes had been described by previous authors. The new effects reported by Ritter were not confirmed by later researchers.

82

83

Alessandro Volta, “On the electricity excited by the mere contact of conducting substances of different kinds,” Philosophical Transactions of the Royal Society 90 (1800), pp. 403–431, at 420–427. Johann Wilhelm Ritter, Versuche und Bemerkungen ueber den Galvanismus der Voltaischen Batterie, Annalen der Physik 7 (1801), pp. 431–484, at 474–475.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

357

10. ØRSTED MEETS RITTER Ørsted first met Ritter when the later was publishing those results. It is well known that in the summer of 1801 Ørsted began a series of travels in Europe. He first visited Germany, where he met Fichte and Schlegel (in Berlin) and later Schelling and Ritter in Jena. After spending several months in Germany he travelled to France and to the Netherlands and returned to Copenhagen in the end of 1803. Ørsted was deeply interested in galvanism—as many other people at that time—and, having heard about Ritter’s researchers, obtained a letter of introduction and met him on the 18th of September. In the following days Ritter showed him many new experiments, and Ørsted was immediately influenced by him. He was soon convinced that Ritter’s work was of the highest importance. A few months later, Ørsted travelled to France, where he began to publicise Ritter’s work. Ritter’s experiments and ideas will be hereafter presented, whenever possible, through Ørsted’s reports.84 The relevance of this kind of source as providing an elucidation of Ørsted’s ideas will be discussed later. Ørsted communicated some of Ritter’s experiments to the Société Philomatique, and obtained a positive reaction. He then wrote to Ritter and asked him to communicate any new discoveries he made, and he was “very flattered” for receiving a series of letters with detailed descriptions of his experiments. Ritter authorised him to announce all his new discoveries to the French physicists.85 From his writings about Ritter’s researches it is possible to perceive that Ørsted espoused his ideas and did not doubt his experimental findings. Ørsted claimed that all previous researchers had given little attention to the effects of electricity upon the organs of sensation. Ritter, however, studied those effects with great care, “even at the price of risking his own health.”86 Mr. Ritter reduced all the effects of the pile on the animal body to expansions and contractions. The positive pole increases the volume of several parts of the human body; and the negative one produces a diminution of the same parts. When the tongue is put into contact with the positive lead, and the negative one is applied to any other part of the body, and they are left in such a position for some minutes, there arises a small boil on the tongue. When the negative lead is put in contact with this organ, it likewise produces a small depression. When the wet hands are put in contact with the poles of the pile for some minutes, the pulse of the hand in contact with the positive pole becomes sensibly stronger, and that of the hand touching the negative one becomes weaker. […] Expansion is followed by a sensation of heat, and contraction by a sensation of cold.

84

85

86

Kirstine Meyer, “The Scientific life and works of H.C. Ørsted,” in Meyer (ed.), H.C. Ørsted, Scientific Papers.…, pp. xxiii–xxiv; Dan Ch. Christiansen, “The Ørsted-Ritter partnership and the birth of Romantic natural philosophy,” Annals of Science 52 (1995), pp. 153–185. Hans Christian Ørsted, Expériences sur un appareil à charger d’électricité par la colonne éléctrique de Volta; par M. Ritter, à Jena. Présentées à l’Institut National par J. C. Ørsted, docteur à l’Université de Copenhagen, Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts 57 (1803), pp. 345–368, at 368. Hans Christian Ørsted, Expériences avec la pile électrique…, p. 404.

358

R. DE A. MARTINS The action of the pile on the organs of the senses is modified by the particular nature of each of them; it is remarkable that the two poles of the pile produce in some way the two extremes of each type of sensation.87

Perhaps Ørsted had not been able to reproduce all those effects—or, maybe, he heard that other people did not confirm them—and added this remark: Those experiments require a lot of care. To repeat them successfully it is necessary to be acquainted with the exhaustive descriptions that the author has provided in several highly detailed treatises.88

Later publications by Ørsted allow us to conclude that in the following years he maintained his acceptance of Ritter’s results, and deemed them of the utmost importance. In 1807, in a review of the most relevant recent advances of chemistry (in his opinion), Ørsted called the attention of his readers to Ritter’s experiments: […] electricity, especially in the form in which it appears in galvanism, is capable of producing the extremes of all sensations; in the gustatory organ acidity and alkalinity, in the olfactory organ a similar contrast, in the eye the two extreme prismatic colours, in the ear higher and deeper tones, in the tactile sense change in temperature and expansion and contraction, in the nerves changed incitability.89

11. RITTER’S SEARCH FOR THE RELATION BETWEEN POLARITIES Both Ritter’s discovery of ultraviolet radiation and his studies of the physiological effects of galvanism show his experimental involvement with the study of polarities. Was this a mere empirical search, or was it guided by some theoretical ideas? Anja Jacobsen has already emphasised that Ritter was attempting to demonstrate, by experiment, his belief about the fundamental principle of polarity in different areas of natural philosophy.90 At the time (1801) when he discovered the invisible “chemical rays” (ultraviolet rays), Ritter announced his research programme in the following words: It will be the result of a larger factual investigation to exhibit the polarity of chemistry, electricity, galvanism, magnetism, heat, etc., in accordance with their principles as one and the same in all.91

Notice two fundamental ideas that clearly appear here: the polarity of all forces and their unity. We can compare Ritter’s words with Schelling’s: “[…] it is the first principle of a philosophical doctrine of nature to go in search of polarity and dualism throughout nature.”92 (Whether Ritter got this idea from Schelling or not

87

88 89 90

91 92

Ibid 404–405. A shorter version of this account can be found in: Ørsted, “Experiments with the electric pile, by Mr. Ritter, of Jena,” Journal of Natural Philosophy, Chemistry, and the Arts, 8 (1804), pp. 176–180. Ibid p. 405. Ørsted, “Reflections on the history of chemistry,” (1807), in Selected Scientific Works…, p. 252. Anja Skaar Jacobsen, Between Naturphilosophie and tradition: Hans Christian Ørsted’s dynamical chemistry (Unpublished Ph.D. thesis, University of Aarhus, 2000), p. 70. Ritter, apud ibid. p. 60. Schelling, On the world soul, apud Schelling, Ideas for a philosophy of nature, translated By Errol E. Harris and Peter Heath (Cambridge: Cambridge University Press, 1995), p. ix.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

359

will be discussed later. Whatever the origin of his ideas, it is clear that Ritter’s main concepts were polarity and the unity of all forces in nature. All phenomena were produced by two opposite forces, actions or qualities, tending to equilibrium. Every force in nature was limited by an opposing force. And Ritter conceived heat, light, electricity, galvanism, and magnetism as different forms of the fundamental forces of attraction and repulsion.93 Johann Wilhelm Ritter (1775–1810) may well have concluded on the basis of Galvani’s discoveries that the same forces which generate electricity also produce chemical effects, but Volta’s final discovery threw far more light on this truth. Ritter used this with rare spirit and power to show how the same natural forces manifest themselves in chemical, electrical, and magnetic effects, in light, in heat, indeed, even in the manifestations of life in organic bodies.94

In later works, Ørsted regarded the identity of electricity, galvanism, chemical forces, magnetism and the “space-filling forces” as the basic idea of the new chemistry, and presented a list of the forerunners of this idea: Priestley, Wilke, Kratzenstein, Herny, Karsten, Forster, Gren, Lichtenberg, Hube.95 According to him, however, none of them had attained the central discovery: However, as these otherwise outstanding men, misled in part by the assumption of a characteristic electrical substance, regarded the particular mode of action which we call electricity as the basis for all other phenomena, instead of assuming it to be one of the various realizations of the universal natural forces, they limited their horizon and gave the entire grand theory, which should have originated from this, the appearance of a narrow hypothesis. On the other hand, it must be admitted that it was scarcely possible to develop this view more completely before our knowledge of electricity and several related effects had progressed to greater maturity so that this advance was naturally reserved for more recent times. Ritter can therefore be regarded as the creator of modern chemistry. His comprehensive ideas and his achievements, undertaken with such great vigour and exertion, spread a great light in all directions. To a certain extent, Winter [sic] deserves to be placed next to him.96

It is essential to distinguish Ritter’s ideas from our current views on energy conservation and the possibility of transformation of the different “forces” in one another. Ritter was not attempting to find a quantitative relation between the amount of different forces that could be mutually transformed. He accepted that natural phenomena are produced by a pair of opposite fundamental forces, and that those contrary powers will become manifest as different polarities in diverse conditions. Since all polarities have a common source, they must exhibit specific, non-arbitrary connections. In a paper published in 1803 presenting “A review of the latest advances in physics,” Ørsted referred to the associations that had been found by Ritter and attempted to provide a rationale for some of them: This student of nature [Ritter] has also brought us enlightenment concerning the other senses, but so far it has yielded less conspicuously satisfying results. The fact

93 94 95

96

Jacobsen, Between Naturphilosophie and tradition…, pp. 60–76. Ørsted, “First introduction to general physics,” (1811), in Selected scientific works…, p. 304. Ørsted, “View of the chemical laws of nature obtained through recent discoveries,” (1812), in Selected scientific works…, p. 312. Ibid. pp. 312–313; my emphasis.

360

R. DE A. MARTINS that the positive pole leaves an acid, the negative an alkaline taste on the tongue had previously been noticed, but that the positively galvanized ear perceives sounds as softer, the negatively galvanized as louder, is a discovery of the same nature. The negative pole evokes an ammoniac smell, the positive seems to deaden the smell. As far as touch is concerned, galvanism has a hot and a cold pole, about which we may soon expect more detailed enlightenment. If we now take a look at previously known facts from this vantage, we see once more a great many phenomena converging towards a focal point. The positive pole generates oxygen gas from water, and this transforms combustible substances into acids or acid-like substances; the negative generates hydrogen gas, and this is a primary component of the few alkalis which we have so far been able to decompose. This yields enlightenment about the effect of galvanism on both taste and smell, but in the latter regard, we need to note that oxidizing substances (like gas muriatique oxygéné) also suppress smell and cause catharrs. As far as hearing is concerned, we recall that, according to Chladny’s [sic] discovery, the notes of a flute sound far higher in hydrogen gas than in oxygen gas.97

Notice that, for Ørsted, the relations between the poles of the pile and the specific sensations they produce are not arbitrary, but should be understandable in a broader framework. Above all physiological effects of galvanism, Ørsted emphasised the relation between the positive pole and expansion, and between the negative pole and contraction: However, it will forever remain a major discovery concerning the effect of galvanism that the positive pole causes expansion, the negative contraction. This law, in its nature so simple, in its application so fruitful, already explains why the eye sees everything larger in the positive state and everything smaller in the negative. At this moment, it would be too daring to establish all the important conclusions which can be drawn from this discovery. Instead, we want to recall only two secondary discoveries by the same student of nature [Ritter] which give cause for much thought. If the tongue is positively galvanized (of course continuously), a swelling appears at the affected spot, whereas a depression is produced by the negative pole. Positive galvanism makes the pulse big, negative makes it small. (Here, fast and slow should not be confused with big and small.).98

Let us recall that in his first communication of Ritter’s physiological researches he had already pointed out this important feature: “Mr. Ritter reduced all the effects of the pile on the animal body to expansions and contractions.”99

12. THE TWO FUNDAMENTAL FORCES OF NATURE According to Schelling’s Naturphilosophie, the productive nature has two opposite activities: repulsion and attraction, which provide the basis for all polarities in nature. Phenomena can take place only when there are oppositions. All natural effects are the products of two opposing powers: “[…] Nature is able to achieve the entire manifold of her phenomena, on the small scale as well as on the large, by means of opposing forces of attraction and repulsion.” Positive force is a repulsing,

97 98 99

Ørsted, “A review of the latest advances in physics,” (1803), in Selected scientific works.…, p. 108. Ibid. p. 108. Ørsted, Expériences sur la pile électrique…, p. 404.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

361

expansive, elastic or repelling force. Negative force is restrictive, attractive.100 The interplay of those opposite forces produces all natural phenomena: In order to maintain this perpetual exchange, Nature had everywhere to count upon contradictories, had to set up extremes, within which alone the endless multiplicity of her phenomena was possible.101

The same idea appears in Ritter’s works: Proof for the absolute polarity in nature. Nature is activity [Handeln], and it is nature only to that extent. Activity requires however a diversity, because only by this does activity arise[…].Each action thus presupposes difference. This however is contrast, polarity. And nature only is, where activity is, therefore polarity must be everywhere.102

The conflict and mutual replacement between opposite forces was regarded as the source of all phenomena. The opposite electrical charges, the two magnetic poles, the contrast between acids and bases and many other dualities were regarded by the Romantic philosophers as examples of this basic polarity of nature. According to Schelling, matter can be reduced to the fundamental forces: “Matter and bodies, therefore, are themselves nothing but products of opposing forces, or rather, are themselves nothing else but these forces.”103 This entails a fundamental unity of all kinds of matter: “All matter is intrinsically one, by nature pure identity; all difference comes solely from the form and is therefore merely ideal and quantitative”104; “[…]everything we call matter is simply a modification of one and the same matter, which admittedly, in its absolute state of equilibrium, we do not know by sense, and which must enter into special relationships to be knowable for us in this way.”105 In the same way as there is only one single basic matter, according to Schelling there is one single pair of opposite forces that can display different forms. Schelling attempted to identify positive and negative electricity respectively with the fundamental repulsive and attractive forces by taking into account their general properties: We can accordingly state the general law of the electrical relation of bodies thus: That one of the two [bodies] which enhances its cohesion in opposition to the other will have to appear as negatively electric, and that one which diminishes its cohesion, positively electric.106

Instead of multiplying forces to explain the variety of natural phenomena, Schelling searched for a hidden unity: But our mind strives towards unity in the system of its knowledge. It does not tolerate a special principle being thrust upon it for every single phenomenon, and it believes that it sees Nature only where it discovers the greatest simplicity of laws amid the greatest variety of phenomena, and the most stringent parsimony of means in the highest prodigality of effects.107 100 101 102

103 104 105 106 107

Schelling, Ideas for a philosophy of nature…, pp. 135, 187. Ibid. p. 87. Johann Wilhelm Ritter, Fragmente aus dem Nachlass eines jungen Physikers (Leipzig: Infel, Verlag, 1938), p. 31. Schelling, Ideas for a philosophy of nature…, p. 156. Ibid. p. 137. Ibid. p. 223. Ibid. p. 118. Ibid. p. 111.

362

R. DE A. MARTINS

The interplay of the two basic forces produce all qualitative changes in matter, and for that reason Schelling regarded their study as the foundation of chemistry: “The subject-matter of chemistry is attractions and repulsions, combinations and separations, insofar as they depend upon qualitative properties of matter.”108 The same attractive and repulsive forces that are the basis of electricity are also the causes of chemical affinity: That which makes substances negatively electric is at the same time that which makes them combustible, or, in other words: Of two substances, that which has the greatest affinity for oxygen always becomes negatively electrified.109 We can find similar ideas in Ørsted’s writings: […]all chemical effects can be traced back to the manifestation of two principal forces, widespread throughout nature, whose properties in their free state, however, cannot easily be found by chemical means. From another side, however, we have arrived at greater knowledge of these forces. In electric, galvanic, and magnetic effects two opposite forces have been found, widespread throughout nature, and it has been possible to investigate the laws which govern their freest manifestations and pursue them through the most diverse conditions to the point where they also produce chemical effects.110 The dynamic theory […] extends the scope of chemistry far beyond its old bounds. Electricity, magnetism, and galvanism now become part of chemistry, and it is shown that the very same fundamental forces which generate these effects also produce the chemical ones in another form.111

One usually ascribes the origin of this dynamical view of nature to Immanuel Kant.112 In some sense this is true, since Kant reduced the basic properties of matter to attractive and repulsive forces. However, Kant did not attempt to include in his dynamical view the different forces that concerned the Romantic philosophers: light, heat, electricity, magnetism, galvanism, and chemical forces. It was Friedrich Schelling who took this step. Kant’s opposite forces had a simple function: to account for the structure of matter (and, perhaps, its density). Schelling’s polarity, on the other hand, accounted for several phenomena such as magnetism, electricity, and chemical forces, being therefore widely different from Kant’s attraction and repulsion.113 According to Robert Stern, it was Fichte (not Kant) who was the main influence acting upon Schelling’s views on the basic forces of matter.114 Nowadays, as Kant is a more honorable ancestor than Schelling, several authors attempt to detach the Romantic nature philosophy from Schelling and connect it exclusively with Kant. That seems to me a mistake, however.

108 109 110 111 112

113 114

Ibid. p. 206. Schelling, p. 102. Ørsted, “First introduction to general physics,” (1811), in Selected scientific works…, p. 291. Ørsted, “Reflections on the history of chemistry,” (1807), in Selected scientific works…, p. 252. Timothy Shanahan, “Kant, Naturphilosophie, and Ørsted’s discovery of electromagnetism: a reassessment,” Studies in the History and Philosophy of Science 20 (1989), pp. 287–305; Kaiser, “Symmetries in Romantic physics…, p. 78. Jacobsen, Between Naturphilosophie and Tradition…, pp. 84–85. Stern, in Schelling, Ideas for a philosophy of nature…, pp. xvi–xx.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

363

There are, of course, both similarities and differences between Ritter’s and Schelling’s ideas about the unity and polarity of forces. Although there are differences between their views,115 Ritter always retained the ideas of a unity of the forces of nature, the relation between electricity, chemistry, and other forces, and the principle of polarity, going as far as stating that “there must be polarity everywhere.”116 If one believes that all forces of nature are different forms of a single primary force (Urkraft), he / she will expect to find a deep relation between the poles (or extremes) of the several types of force. Schelling provided several examples taken from the recent scientific findings—especially in chemistry and galvanism—but did not pursue empirical investigations. He believed that speculation was a safe method.117 Ritter did not agree. He added a strongly experimental approach to Schelling’s speculations, and pursued empirical inquiries. Polarity and the unity of forces were the guiding ideas of his researches, but he felt the need to manipulate nature and to observe the results. So, the concept of polarity did not represent for Ritter just a philosophical framework, but also a heuristic principle, directing him to find new phenomena.118 Schelling’s theoretical influence upon Ritter can be noticed in his presentation of the natural phenomena in pairs of polar opposites. On the other hand, Ritter’s experimental findings influenced Schelling.119

13. THE EMBLEMATIC WAY OF THINKING It is important to stress that Schelling’s approach was not just a different philosophical system, but a different way of thinking about nature. To exhibit this difference as clearly as possible, I will choose an indirect route, including a short visit to China in order to elucidate the emblematic way of thinking. The meaning of “emblematic” I would like to apply is not unlike Ashworth’s use of this term, as applied to Renaissance thought. The emblematic world view is, in my opinion, the single most important factor in determining late Renaissance attitudes towards the natural world, and the contents of their treatises about it. The essence of this view is the belief that every kind of thing in the cosmos has myriad hidden meanings and that knowledge consists of an attempt to comprehend as many of these as possible. To know the peaco*ck, as Gesner wanted to know it, one must know not only what the peaco*ck looks like but what its name means, in every language; what kind of proverbial associations it has; what it symbolizes to both pagans and Christians; what other animals it has sympathies or affinities with; and any other possible connection it might have with stars, plants, minerals, numbers, coins, or whatever.120 115

116 117

118 119 120

Ritter, Goethe and Novalis criticized Schelling’s speculations, but they adhered to the idea of polarity. See Kaiser, “Symmetries in Romantic physics…,” p. 80. Hermann, “Unity and metamorphosis of forces…,” p. 58. After the discovery of electromagnetism and electromagnetic induction, Schelling claimed that the relations between chemistry, magnetism and electricity had been anticipated by German philosophers—including himself, of course. See Hermann, “Unity and metamorphosis of forces…, p. 60. Kaiser, “Symmetries in Romantic physics…, p. 81. Jacobsen, Between Naturphilosophie and tradition…, pp. 66–67. William B. Ashworth, Jr., “Natural history and the emblematic world view,” in David C. Lindberg and Robert S. Westman (eds.), Reappraisals of the Scientific Revolution (Cambridge: Cambridge University Press, 1990), pp. 303–332, at 312.

364

R. DE A. MARTINS

When Marcel Granet attempted to describe ancient Chinese thought to 20th-century occidental readers, he stressed that the Chinese words did not correspond to concepts. They were not mere signs, but should be understood as emblems.121 Instead of corresponding to general and abstract ideas, the words evoke “an indefinite complex of particular images.” A particular emblem may encompass what we would describe as incompatible concepts related to space, time, colour, etc. The same emblem, for instance, characterises the era and empire of the Tchou dynasty, the red colour, the summer season, the South region. The East region is related to benevolence, flexibility, to muscles, to the liver, to the spring season and to green colour.122 According to Granet, the emblematic character of the Chinese words establishes a series of relations that would be meaningless if one attempted to understand them as conceptual links: Mountains and humpback people are abundant to the West and they characterise it, in the same way as the heaps of the harvest that evoke Autumn. A hump is a skin excrescence; the skin depends on the lungs, lungs depend on the Autumn and are related to the white colour. But when we refer to skin we refer to leather and armour, that is, war and punishment. So, the western barbarians are regarded as endowed with a warlike humour, and executions—both military and penal ones—are reserved to the Autumn, and the Spirit of Punishment, who is remarkable by his white hair, lives in the West. Hair comes from the skin, and white is the meaningful emblem of West and Autumn, and also of the Yin age. That era was inaugurated and characterised by the kingdom of T’ang the Victorious, a hero who became famous for the punishments he inflicted and because of his habit of walking with his body completely bent.123

This way of thinking is alien to contemporary scientific thought, but it is not far from the way of thinking introduced by Schelling and used by Ritter and Ørsted in some of their researches. There are other similarities that can be found between the ancient Chinese thought and Schelling’s fundamental polarities. In one of the chapters of his book Granet discusses the meaning of the couple of words yin and yang. Some interpreters of the Chinese thought construe yin and yang as two forces. Other scholars interpret them as substances. However, those are occidental conceptual categories that do not apply in a strict form to the Chinese thought.124 Yin and yang might be regarded both as forces and substances, and also as corresponding to other categories; however, they can also be regarded as neither forces nor substances. They are general opposite and complementary conditions that follow each other. According to Granet, in an ancient book, the Shi Jing, the word yin evokes the idea of cold and cloudy weather, or rainy sky. It is applied to the inner part of things, to dark and cold places where, during the summer, it is possible to preserve ice. Any shadowy place, such as the north side of a mountain or the south side of a river, is also described as yin. The word yang, on the other hand, is associated to heat and to the Sun. It may be used to describe the male attitude

121 122 123 124

Marcel Granet, La Pensée Chinoise (Paris: La Renassance du Livre, 1934), pp. 37–39 Ibid. p. 87. Ibid. Ibid. pp. 115–116.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

365

of a dancer in action. It applies to spring time, when the heat of the Sun begins to produce its effects and to the tenth month of the year, when winter begins its retreat. This is the month when buildings should begin to be erected. Sunny and bright places, such as the south side of a mountain and the north side of a river, are also described as yang.125 There are many other meanings associated to those words. The following table presents some of them: Yin

Yang

female cold a closed door something hidden the act of entering inside darkness acute sounds light weight rain night winter earth moon water even numbers

male hot an open door something manifest the act of coming out outside brightness grave sounds heavy weight dew day summer sky sun fire odd numbers

One can perceive a similarity between this Chinese way of comparing widely different things, and Ritter’s association between colours, sounds, temperatures, tastes, etc. Of course, I am not claiming that there was an influence of the old emblematic Chinese thought on Ritter’s ideas, but it is impossible to deny a similarity in the search for polarity or dualities and in the attempt to connect them in an integrated unitary view of nature. 14. THE SYMBOLIC DIMENSION OF NATURPHILOSOPHIE Some authors who wrote about the German Romantic movement have already stressed the symbolic outlook of their world view: “[…] Everything is sign and symbol, […] the whole world has ‘merely indicatory or physiognomic significance’. The whole world should be interpreted as a gigantic system of hieroglyphics, as the language of God or the book of nature.”126 The chain of natural laws which through their actions constitute the essence of every object can thus be regarded as a thought of nature or rather an idea of nature. And as all natural laws together form a unity, the entire world is the expression of an infinite, 125

126

Ibid. pp. 117–118. [The Chinese work is known as the Book of Songs or Odes, given in Granet’s French text as the Che King—Ed.] Alexander Gode-Von-Aesch, Natural Science in German romanticism (New York: AMS Press, 1966), p. 228.

366

R. DE A. MARTINS universal Idea which must be one with an infinite Reason, alive and active in everything. In other words, the world is merely the revelation of the combined creative power and reason of the Godhead.127

According to Alexander Gode-von-Aesch, this idea was born in the late 18th century and was developed by Ritter (and other philosophers). This way of thinking led to identify widely different but symbolically related entities: “[…] it might be shown that romantic imagery tends unconsciously to be of the identity type. When Hardenberg says, for instance, that the brain resembles the testes, he is doubtless in the midst of his magic idealism and conceives of thinking as a procreative act.”128 The symbolic way of thinking is clearly exhibited in Schiller’s, Ritter’s, and Ørsted’s writings: Finally, for the ultimate task of a physics of chemistry, which also has to depict in these phenomena the totality alone, it is necessary to grasp their symbolic character and connection with higher relationships, since every body of individual nature is again, in its idea anyway, a universe. Only if we seek among chemical phenomena, no longer for laws that are peculiar to them as such, but for the general harmony and regularity of the universe, will they come under the higher relationships of mathematics […].129 Each point in the universe is a nature en miniature, but in everything the artist copied the original from another side.130 The ancients worshipped the universal substance under the name of Vesta (Hestia), and this indeed is the sensible image of fire. In this they left us a hint that fire is nothing other than the pure substance breaking through in corporeality, or a third dimension […].131 The most perfect process of combustion will display itself to us where the conflict of universal and particular is perfectly equalized in that attempted process of generation, where the universal and particular of relative cohesion reaches indifference, yielding the hermaphroditic product of water, which, as absolute liquid, is not only the total extinction of the first two dimensions in the third, but also, through the particular is wholly Earth and through the universal wholly Sun; and just here in this equalization the Sun breaks through most completely, except that because of the element of Earth which is included therein, it cannot show itself purely as light, but only as fire (light combined with heat).132 White is the color […] which keeps the eye healthy; the light of the sun is white[…]white presents purity, innocence, love, harmony, etc. […] Also the water is white, harmony, purity, innocence, the source of everything on Earth.133

In his Reflections on the History of Chemistry (1807) Ørsted remarked that the medieval thought and modern science were widely different, but had some features in common, including the search for unity:

127 128 129 130 131 132 133

Ørsted, “First introduction to general physics,” (1811), in Selected scientific works…, p. 252. Gode-Von-Aesch, Natural Science in German romanticism…, p. 219 Schelling, Ideas for a philosophy of nature…, p. 220; my emphasis. Ritter, Fragmente aus dem Nachlass …, p. 57. Schelling, Ideas for a philosophy of nature…, p. 65. Ibid. p. 66; my emphasis. Ritter, Fragmente aus dem Nachlass…, p. 39.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

367

The mystical tendency of the Middle Ages is so opposite the striving of our time towards perfect clarity that it might easily seem impossible for them both to have a share in the truth. To deny the contrast between them would be against evident truth, but no contrast can exist where there is nothing in common.[…] Every effort towards insight into nature aims at bringing separate phenomena under a common terminology, at discovering laws which everything obeys, in short, at bringing the unity of reason to nature. The mystical age had at least this endeavour in common with ours.134

Ørsted understood this search for unity as symbolic. As an instance, he discussed the medieval belief in a relation between the planets and the metals: “At first glance, this seems mere fantasy, but if we consider the matter more closely, we find an underlying truth.” He then presented some arguments favorable to the idea, such as this: “[…] it is worth noticing that gold, which was the sun of metals according to that time, is deposited primarily around the equator and also maintains its metallic nature most perfectly in all assays.” Finally, he acknowledged that there was (yet) insufficient basis for such a comparison: “I admit that, in spite of all our greater knowledge, we are still not able to advance such a comparison between the metals and the planets, but the basic idea is hardly to be disdained.”135 In his own work, Ørsted usually employed this symbolic method. One instance is his view of positive and negative electricity: The primary form of positive electricity is the radiating point, of the negative, on the other hand, the circle so that one seems to form the internal, the other the external, one the point which radiates from its centre in all directions, the other the enclosing periphery. The natural symbol of electricity, then, is a circle with its radii, the symbol of positive electricity the radiating point, and of negative electricity the point surrounded by concentric circles. These symbols undoubtedly deserve our fullest attention, for they reappear everywhere, and who knows whether all of Nature’s mathematics does not lie hidden in them! (ØRSTED, On the harmony between electrical figures and organic forms, 1805, in Selected scientific works, p. 185)

In 1804 Ørsted described that the two galvanic poles produced different forms similar to a vegetation and its roots: […] I put a solution of acetate of lead in contact with the poles of the pile. The dissolved lead calx should be oxidized more strongly on the oxygen side and be precipitated as brown lead calx, but on the hydrogen side it should be reduced and thus be precipitated. This indeed happened. On the hydrogen side I obtained a beautiful metallic lead vegetation but on the oxygen side a brown lead calx formed shapes comparable to the positive soot figures. I would prefer to compare these shapes with plant roots. Could it be that oxidation and deoxidation are associated with definite forms which occur if no external causes oppose them? Could the organic forms be necessary products of the internal chemical process?136

In 1805 Ørsted applied this idea to the structure of trees: […] In the simplest, purest experiments, which actually serve as a basis for all the chemical discoveries of more recent physics, we find the process of reduction (deoxidation) united with the external form of vegetation, whereas the process of combustion is accompanied by a form whose boundary is the circumference of a circle when it radiates from a central point or parallel lines when it radiates from a central line, that is, we see in it the norm of the internal form of a plant. Therefore we should

134 135 136

Ørsted, “Reflections on the history of chemistry,” (1807), in Selected scientific works…, p. 247. Ibid. p. 248. Ørsted, “Galvano-chemical observations,” (1804), in Selected scientific works, p. 168.

368

R. DE A. MARTINS

Fig. 1. According to Lichtenberg, positive electricity produces radial figures and negative electricity produces concentric circles. Ørsted joined both figures to produce “the natural symbol of electricity,” that might be depicted as above.137

expect to find the same formations everywhere in nature, assuming that the same form must follow the same force unless the effect of foreign forces change it. We need only glance at nature to find our assertion confirmed. The plant lives solely by the influence of sunlight, and thereby it constantly generates oxygen gas and is deoxidized or reduced. The same form and the same chemical process which were united in the electrical effect are so here, too. Internally, however, the plant must oxidize.[…]Another reason can be found in the nature of the plant juices themselves. These are acidic, and those that are not to any noticeable extent still have a tendency in this direction so that they are always acidified by fermentation. Thus we discover the same agreement between form and force in the interior of the plant as in its exterior, and in both the most perfect similarity to what we have seen in electricity. We could add that plant fibers appear as parallels only when viewed from one direction, that is, lengthwise, but when viewed crosswise, the circle is the predominant figure and forces us to acknowledge the negative in the interior of the plant, in every direction.138

137

138

Figure reproduced from Everett Lee and C. M. Poust, “Measurement of surge voltages on transmission lines due to lightning,” General Electric Review 30 (1927), pp. 135–145, on p. 134. Ørsted, “On the harmony between electrical figures and organic forms,” (1805), in Selected scientific works…, p. 186.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

369

15. RITTER’S EMBLEMATIC THOUGHT Ritter’s attention was called by opposite polarities in several phenomena (oxidation versus reduction, hydrogen versus oxygen, etc.) and he attempted to establish definite relations between dualities belonging to different realms. He knew that positive and negative galvanism produced acid and bitter tastes, and he expected that the same causes would produce opposite sensations when applied to the other sense organs. And that was exactly what he thought he had found, as described above. He found out that positive galvanism was related to the red “pole” of light, and negative galvanism to the blue “pole,” and other similar associations that can be ascribed to an emblematic way of thinking.139 The same fundamental polarity was producing different sensations, as it acted upon different organs, as suggested by Schelling: Light and heat are mere expressions of our feeling, not a designation of that which acts upon us. From the very fact that light and heat affect quite different senses, and work so utterly differently, we can already infer that in both cases we are designating mere modifications of our organ.140

One might think that, when Ritter attempted to find a definite relation between such different things as colours and electricity, he was just using some kind of analogy—as Newton’s analogy between the colours of the rainbow and the musical notes. Anja Jacobsen, for instance, explained Ørsted’s parallels between electricity, combustibility, and acidity–alkalinity as the use of the same model of explanation for several phenomena “by virtue of analogies.”141 Andrew Wilson has also pointed out that Schelling, Steffens, Ritter, Winterl, and Ørsted had identified “whole series of analogies between physical phenomena.”142 I will claim, however, that Ritter’s (and also Ørsted’s) emblematic or symbolic thought was something much stronger than a mere analogy. From a historical point of view, the concept of “analogy” was born in mathematics where it meant an equality between ratios or proportions.143 Afterward this word came to be used in several different senses.144 Although there is a wide range of “analogy concepts,” let us assume the following statement as a reasonable account of most recent uses of this word: two objects A and B of any kind are analogous if there are parts, properties or relations that are similar or equal in both A and B (that is, if they have some equivalent features) and if, beside that, they have some difference.

139

140 141 142 143

144

“It seems that Ritter’s attempt is an amazing example for the notion of polarity and how far the idea of polarity could influence experimental data” See Kaiser, Symmetries in Romantic physics…, p. 82. Schelling, Ideas for a philosophy of nature…, p. 134. Jacobsen, Between Naturphilosophie and tradition…, p. 133. Andrew D. Wilson, “Introduction,” in Selected scientific works…, pp. xv–xl, at xxix. Even in ancient Greek thought, analogy was also regarded as a method of suggesting explanations of natural phenomena. G. E. R. Lloyd, “Analogy in early Greek thought,” in Philip P. Wiener (ed.), Dictionary of the history of ideas: studies of selected pivotal ideas (New York: Charles Scribner’s Sons, 1973), vol. 1, pp. 60–63, at 60. Mary Hesse, “Models and analogy in science,” in Paul Edwards (ed.), The Encyclopedia of Philosophy (New York: Macmillan & The Free Press, 1967), vol. 5, pp. 354–359.

370

R. DE A. MARTINS

Notice that, according to this concept, two identical objects are not analogous. If two objects are analogous, this similarity suggests that they might have other equivalent features: The examination of likeness is useful with a view both to inductive arguments and to hypothetical reasonings, and also with a view to the rendering of definitions.[…]It is useful for hypothetical reasonings because it is a general opinion that among similars what is true of one is true also of the rest. If, then, with regard to any of them we are well supplied with matter for discussion, we shall secure a preliminary admission that however it is in these cases, so it is also in the case before us […]145

“Reasoning by analogy” means inferring an unknown similarity between two objects, from a known analogy between them. Of course, reasoning by analogy is not demonstrative. It might produce hypotheses, or conjectures, and have a useful role in research, but cannot lead to certainty. Let us consider one famous instance. In 1895 Wilhelm Conrad Röntgen was studying electric discharges in vacuum tubes, when he noticed that a nearby fluorescent plate became bright. The unexpected phenomenon called his attention, and its study led to the discovery of a new kind of invisible penetrating radiation, with peculiar properties. From the very beginning of Röntgen’s investigation, it became clear that the new radiation could not be explained by existing theories—it was a puzzle, and that was the reason why it was called “X rays.” Röntgen knew it was not light, but his early research on X rays was guided by the search for analogies between the new radiation and light. He checked if it would suffer reflection, refraction, polarisation, diffraction, etc. But he did not assume that the new rays had to exhibit those phenomena.146 The kind of thinking behind Ritter’s experiments is not analogical thought, in the sense described above. His philosophical presuppositions told him that there should be definite correspondences between the poles of all forces in nature. The specific relationship could be suggested by general philosophical considerations, but in most cases had to be discovered empirically. In any case it was certainly there, waiting to be discovered. In Ritter’s mind there was no doubt that there should be a relationship between the electric and magnetic poles and the polarities of oxidation–reduction, red–blue, warm–cold, positive–negative, contraction–expansion, etc.—because all forces of nature arise from the Urkraft and are, in some sense, the same thing. This hidden unity is beyond the reach of our experience, but through the emblematic way of thought it is possible to capture its meaning. The essential polarity of nature cannot be reduced to any of the polarities we observe, but it is possible to have a glimpse of its meaning as the common source of all interrelated dualities we observe. We have now arrived at a point where we recognize the principles of acidity and alkalinity as principles of electricity. These principles are to be found in all bodies and cannot be separated from their nature. We will certainly not claim on this account that all bodies are acids or bases, for it depends not only on whether these principles

145 146

Aristotle, Topics, book I, ch. 18, 108b, 6–16. Roberto de Andrade Martins, Jevons e o papel da analogia na arte da descoberta experimental: o caso da descoberta do raios X e sua investigacão pré-téorica, Episteme. Filosophia e História das Ciências em Revista 3 (6) (1998), p. 222–249.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

371

are present, but also on how they are present. Otherwise, we would be obliged to claim that even the coloured rays of light were acidic or basic. Now this, as a paradox, would not frighten us, but we would become entangled in a great many difficulties. Instead of calling a body with an excess of the positive principle a base, we could say with equal justice that it was violet internally, and that we should not be concerned merely with the outward appearance because there could be causes which impeded the manifestation of the colour. It is indisputable that we should not allow ourselves to be prevented by appearances from seeking the inner principle. Once we have found the principle and, at the same time, seen it revealed in the most varied forms, e.g., as light, as heat, as electricity, as magnetism, etc., it is then time to differentiate precisely between these forms and not to confuse them because of what they have in common.147

The emblematic way of thinking is not always explicitly presented in Ritter’s and Ørsted’s scientific papers, but a careful analysis of some remarks presented by Ørsted will show that it underlies some of the experimental accounts. When Ørsted described Ritter’s discovery of the invisible radiation at the violet end of the visible spectrum, he remarked: Those experiments can be easily applied to some others, made by the same physicist. He kept his eye in contact for a few minutes with the negative lead of Volta’s electric pile, and after this operation all the objects seemed to him red; but after keeping in contact with the positive lead, he saw everything blue.148

Notice that Ørsted is establishing a relationship (not an analogy) between widely different classes of phenomena, according to contemporary science: the colours produced by decomposing white light with a prism, and the subjective colours produced by electric stimulation of the eye. It is also remarkable that Ørsted, following Ritter did not conclude that negative electricity was related to red, and positive electricity to blue, but the opposite: This great discovery was soon joined by a second, that of the effect of galvanism on the eye. If the nerves of the eye have been put into the positive state, all objects are seen with a red color (in darkness) and larger than they are otherwise seen, but if they have been put into the negative state, all objects appear blue and smaller than usual. If we recall that the positive pole of the battery is the oxidizing one, the negative the deoxidizing one, and that the blue color lies closest to the violet in the spectrum, the connection between this and the previous discovery becomes very clear to us. Oxidation and the red pole of the spectrum, deoxidation and the violet pole are associated with each other.149

In a footnote, Ørsted explained that, actually, it was necessary to put the negative end of the pile into contact with the eye to produce the sensation of red colour, but that in this case one should not explain the effect as due to negative galvanism, but as due to the positive galvanism acquired by the retina and optical nerves: Actually, [all objects are seen with a red colour] if the negative pole has been kept in contact with the eyeball for some time. The liquid in the eye, like any other liquid, must polarize, and therefore, if it becomes negative on the outer surface, the inner becomes positive. This explanation stems from the astute Dr. Reinhold in Leipzig, who has also repeated Ritter’s experiments and found them completely confirmed.150

147 148 149 150

Ørsted, “The series of acids and bases,” (1806), in Selected scientific works…, p. 239. Ørsted, Expériences avec la pile électrique…, p. 410. Ørsted, “A review of the latest advances in physics,” (1803), in Selected scientific works…, pp. 107–108. Ibid. p. 107.

372

R. DE A. MARTINS

A similar explanation appeared in his French paper: One should remark that when the outside of the eye is in a negative state, the retina and the optic nerve become positive, and vice versa; because the eye is full of a fluid, in which there occurs the same distribution of electricity that happens in water and other fluids. Therefore, it is in a positive state that the optic nerve perceives all objects with a red color, and in the negative state they appear with a violet color.151

Only accepting this interpretation was it possible to establish a coherent relationship between electricity, chemical effects and color: both positive electricity and the red light produce oxygenation, and both negative electricity and violet light produce reduction (Ørsted, 1803b, p. 410). The careful reinterpretation of the experimental situation was required because Ritter and Ørsted were not describing mere analogies but were trying to unravel the inner correspondences between different manifestations of the Urkraft.

16. THE ELECTRIC POLARITY OF THE EARTH Now, if we return to Ritter’s researches on magnetochemistry and their context, as described by Ørsted, it will become clear that his steps were guided by the above described emblematic way of thinking. Ørsted’s first communication to the French Academy was a report on Ritter’s “secondary pile.” Ritter found out that it was possible to build an electric accumulator using a pile made of a single metal. He built it with a series of metallic plates intermingled with paper wet with salt water. After this secondary pile had been connected to a Voltaic pile for some time, it became a source of electricity. This was a very interesting finding because Ritter had been able to induce an electrical polarity upon a system that was completely symmetrical. This discovery was well received by the French savants, and Ørsted published his report at the Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts.152 However, when he was preparing this communication he received a new letter from Ritter telling him about his fresh discovery of the “electric poles” of the Earth, and Ørsted included this information as a post-scriptum to his paper. During 1803 Ritter had published two papers where he discussed a possible relation between some atmospheric phenomena (including storms and aurora borealis) and electricity or magnetism.153 He pointed out that there were periodic variations of several phenomena that could establish a relationship between them. 151 152

153

Ørsted, Expériences sur la pile électrique.., p. 410. Ørsted, Expériences sur un appareil…. A shorter version of Ørsted paper was translated in: Ørsted, “Abstract of a memoir on galvanism, sent to the National Institute by Mr. Ritter, of Jena,” Journal of Natural Philosophy, Chemistry, and the Arts, 7 (1804), pp. 288–291. Johann Wilhelm Ritter, Auszuege aus Briefen verschiednen Inhalts an der Herausgeber. 1. Von Herrn J. W. Ritter, Annalen der Physik 15 (1803), pp. 106–110; Ritter, Einiges ueber Nordlichter und deren Periode, und ueber den Zusammenhang des Nordlichts mit dem Magnetismus, und des Magnetismus mit den Feuerkugeln, dem Blitze und der Electricitaet, Annalen der Physik 15 (1803), pp. 206–226. John Robinson and other authors had already reported that the aurora borealis acted upon the magnetic compass, deviating it from the meridian. See Mottelay, Bibliographical History of Electricity and Magnetism…, p. 309.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

373

In those studies about atmospheric phenomena Ritter had already hinted that the Earth should have an electrical polarity. Then, using his secondary pile (or accumulator), he noticed that the device exhibited weak effects even when it had never been linked to a Voltaic pile. Those effects could be observed using a frog as a sensor. When the secondary pile was put in a vertical position, the upper end of the accumulator acquired a positive charge, and the lower one a negative one. He supposed that this effect was due to an external electrical field produced by the Earth, and moved the secondary pile to several different positions, to find out the direction of the field. Keeping the device in the plane of the magnetic meridian, the effect was maximum when the pile was tilted to the North, and formed an angle of about 30 degrees with the vertical direction. When the secondary pile was put in the horizontal position, in the North–South direction, the North end acquired a positive charge. The effect increased when this extremity of the device was turned about 30 degrees to the East. His conclusion was that the Earth has electric poles and electric meridians. According to Ritter, those poles affect atmospheric phenomena (such as storms) and they produce an electrical polarity in animals, plants, men, stones and all objects.154 In his following letter to Ørsted, Ritter described new experiments using a secondary pile made of 1,000 plates. The device was about 4 meters long and it was difficult to manipulate. The experiments had to be done outdoors, and of course it was very difficult to produce repeatable results with frogs in those conditions. Ritter also told Ørsted that he had been successful in building something that could be described as an electric compass, that pointed towards the electric poles of the Earth. He took a thin gold wire and connected its ends through moist conductors to a 200-elements voltaic pile.155 After five minutes the gold wire was put on a pivot similar to those of magnetic compasses and was protected from air drafts. According to Ritter, the gold needle turned to the electric poles of the Earth.156 Ritter’s experiments with the gold wire were witnessed by Christian Bernoulli, who published a positive report about them.157 Ørsted usually attempted to replicate Ritter’s experiments. In the specific case of the electric compass, he repeated it using a platinum wire, but the experiment did not succeed. He commented: “I would not dare to doubt Mr. Ritter’s experiment because of that; I have repeated it without being completely aware of its details.”158 It is rather curious that in later experiments Ritter built a lengthy (six inches long) bimetallic needle (half its length made of zinc and the other half made of silver) and described that this needle behaved as a magnet, the zinc end pointing

154 155

156

157

158

Ørsted, Expériences sur un appareil…, pp. 364–365. Ibid. p. 365. Ritter had noticed that it was possible to produce an electrical polarity upon metals by this method. Ritter published his first claim concerning this effect in Auszuege aus Briefen…, but he did not provide a description of his experiments. His account was published in 1805: Ritter, Das electrischen System der Körper (Leipzig: C. H. Reclam, 1805), pp. 383–384. Christian Bernoulli, “New galvanic experiments by M. Ritter. Extracted from a letter from M. Christ. Bernoulli,” Philosophical Magazine 23 (1806), 51–54. Ørsted, Expériences sur un appareil …, pp. 364–365.

374

R. DE A. MARTINS

to the North and the silver end pointing to the South. Besides that, the needle was also acted on a magnet, in the same way as a magnetic needle.159 Ritter’s 1803 experiment inspired Jean Nicholas Pierre Hachette and Charles Bernard Desormes to attempt an interesting experiment. In 1805 they built a huge copper-zinc pile containing 1,400 metal plates. The length of the pile was about 1 metre. It was put in a horizontal position in a small wooden boat, floating in still water. They expected the boat to turn to the electrical poles of the Earth, but no motion was observed.160 In the following years (1804–1806) Ritter continued to compare magnetism to electricity. He published a book where he described new evidence for the electric poles of the Earth he had discovered in August 1803. He used several needles made of gold, silver or copper submitted for a few minutes to the voltaic pile.161 One of the extremities of the needles pointed towards some direction between north-north-west and north-west. Ritter described new relations between electricity and magnetism. He built a compass with a long silver-zinc needle and reported that it would behave as a magnet, aligning itself in the direction of the magnetic meridian. The zinc end approached to the North and the silver end to the South. The north pole of an iron magnet would attract the silver end and would repel the zinc end. Therefore, positive electricity pointed to the North magnetism, and negative electricity corresponded to South magnetism.162 The effect was stronger when Ritter replaced the silver part of the needle with carbon or lead. Therefore, when two different metals (or conductors) are connected, they bring forth a polarity that could produce both magnetic and electric effects. When Ritter’s researches on magnetochemistry are regarded in this context, it is possible to perceive that they were not isolated empirical findings suggested by a loose analogy. They must be considered as part of a research programme guided by strong philosophical presuppositions (unity of all forces of nature, basic polarity of forces and their effects) and an emblematic way of thinking. All this led Ritter to search for definite relations between the magnetic poles and the other polarities of nature—electrical, chemical, etc. Ørsted interpreted those results as a demonstration that magnetism and electricity are produced by the same basic forces: […]the same forces which manifest themselves in electricity also manifest themselves in magnetism, although in another form. Attractions and repulsions are the same in magnetism as in electricity, opposite forces attract, like ones repel each other. Through magnetism two pieces of iron can be made to produce the same effect on a prepared frog as two different metals. If an iron wire is magnetized, the end which becomes the south pole will become more combustible than it was before, but the one that becomes the north pole will lose some of its combustibility. Ritter has convinced us of this

159 160

161 162

Ritter, Das electrische System der Körper…, p. 379. When a weakly magnetised iron bar of the same weight was put in the same boat, it soon acquired the North–South direction. Jean Nicolas Pierre Hachette, “Experience sur le magnetisme de la pile electrique,” Correspondance sur l’Ecole Impériale Polytechnique, a l’usage des éleves de cette Ecole,” 1 (1805), pp. 151–153. See also Hachette’s later account of his experiment: Hachette, Sur les expériences electro-magnétiques de M.M. Ørsted et Ampere, Journal de Physique, de Chimie et d’Histoire Naturelle 91 (1820), pp.161–166, at 165. Ritter, Johann Wilhelm. Das electrische System der Körper…, pp. 383–384. Ibid. 379–380.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

375

through many experiments whose validity can easily be ascertained through experience. Consequently, the same forces are at work in electricity and magnetism.163 The remaining similarities between magnetism and electricity are so great that we need only remove the apparent contradictions in order to accept the identity of the forces in them. […] Ritter has also found that magnetized iron wire is less oxidizable at its northern end and more oxidizable at its southern end than iron, but iron or soft steel must be used here because harder steel produces less activity and, in fact, in the reversed order due to its poorer conduction and its corresponding smaller quantity of force. Under similar conditions, muscular contractions are also induced in a prepared frog if two opposite poles of a magnetized iron wire are connected to it in such a way that a closed circuit can be formed. The wires must be magnetized by means of relatively strong magnets. These experiments are still somewhat disputed by physicists, but so many have been successful that it is not easy to assume a false conclusion. […] Therefore, all the functions which can be demonstrated in electricity can also be observed in magnetism: attractions and repulsions, chemical difference, effects on the living animal body, the production of light.164

In 1805, Ørsted used Steffens’ ideas to connect electricity, magnetism, the “four chemical elements” (carbon, nitrogen, oxygen, and hydrogen) and the four principal geographical directions: […] Oxygen and hydrogen, carbon and nitrogen are also here revealed as the 4 chemical elements, the two former corresponding to the contrast in electricity, the two latter to that of magnetism, as our great natural philosopher Steffens first proved. Carbon and nitrogen appear in chemical action, like magnetism in nature, in internally determined forms; oxygen and hydrogen, like electricity, as eternally mutable, striving towards new forms.165

Ørsted remarked that even the “magnetic” pair (carbon and nitrogen) had also an electrical polarity: “The substances containing carbon form the negative, the substances containing nitrogen the positive elements.”166 Carbon and nitrogen would be related to north and south, as shown by geology: “To the north, carbon is prevalent, which is indicated by the enormous number of forests, peat bogs, coal, etc., but to the south, nitrogen is found more often, which is demonstrated by many coral mountains.”167 Next, he presented the complete symbolic relation between electricity, magnetism, the chemical elements, night and day, animal and vegetable, winter and summer, and the four cardinal directions: The day is deoxidizing, the night oxidizing. The same relation reappears on a larger scale between summer and winter. Briefly, a constant process of combustion and reduction proceeds from east to west, the same electro-chemical process which we have demonstrated in the animal and vegetable kingdoms. Steffens’s glorious idea to regard oxygen and hydrogen as representative of east and west, and carbon and nitrogen as representatives of north and south is then confirmed in the most perfect way, however paradoxical it might appear to all those who are not informed about recent physics.168 163

164

165

166 167 168

Ørsted, “New investigations into the question: What is chemistry?,” (1805), in Selected scientific works …, p. 196. Ørsted, “View of the chemical laws of nature obtained through recent discoveries,” (1812), in Selected scientific works…, p. 379. Ørsted, “On the harmony between electrical figures and organic forms,” (1805), in Selected scientific works…, p. 189. Ibid. Ibid. p. 190. Ibid.

376

R. DE A. MARTINS

Although Ørsted did not associate Steffens’ ideas to Ritter’s experiments, one may notice the agreement between their conclusions concerning the East–West electrical polarity of the Earth.

17. ØRSTED AND NATURPHILOSOPHIE The previous sections attempted to show that Ritter’s researches on the polarities of electricity, magnetism, and other forces, can be regarded as a clear example of an empirical quest guided by the assumptions and way of thinking promoted by Schelling’s Naturphilosophie. Ørsted’s uncritical dissemination of Ritter’s ideas and results seems to point out that he accepted all those beliefs in his early scientific career. At this time, Ørsted was presenting to the German-speaking public Winterl’s chemistry. As Kenneth Caneva has convincingly shown, Ørsted modified Winterl’s ideas so as to fit his own beliefs169: “To a very considerable extent, the Winterl who has come down to us is the Ørstedized version closely associated with the dynamical Naturphilosophie Winterl himself stood apart from.” It is reasonable to assume that, at the same time, when Ørsted presented Ritter’s ideas, he would change and adapt them if they did not fit his own beliefs. Hence, I assume that whenever Ørsted is describing Ritter’s ideas and experiments in the early years of the 19th century, he is describing what he accepts as true. Anja Jacobsen has already stressed that it is difficult to distinguish between Ritter’s and Schelling’s influences upon Ørsted: It is quite difficult to distinguish precisely which influence on Ørsted’s ideas stems from Ritter’s electrochemistry and which from Schelling’s Naturphilosophie. Historians of science generally seem to be unclear about how Ritter’s and Schelling’s ideas stand in relation to each other; who influenced the other? However, it is a fact that Ritter’s ideas are more tangible and related to actual experiments whereas Schelling’s ideas are on a more philosophical framing level, although they are sometimes quite similar to each other.170

Ørsted rejected the work of purely speculative philosophers who were not acquainted with experimental work.171 In this respect, Schelling’s work did not seem to him adequately scientific. This does not entail, however, that he was not influenced by Schelling. Although several authors (including Andrew Wilson) have already presented clear substantiation concerning the relation between Ørsted and Schelling, let me add some more evidence. In his 1799 work on Fundamentals of the metaphysics of nature, Ørsted followed Kant in his introduction of the basic forces of matter (attraction and repulsion) as necessary conditions of the existence of matter of finite size: The expansive force prevents the attractive force from reducing the extent of matter to zero, and the attractive force prevents the expansive force from giving matter an 169 170 171

Caneva, “Ørsted’s presentation of others’—and his own—work,” (this volume). Jacobsen, Between Naturphilosophie and tradition…, p. 47. Ibid. p. 71–73.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

377

infinitely large extent. They work in opposition to each other and produce motion in opposite directions so that one may be regarded as negative when the other is regarded as positive.172

Notice that, here, Ørsted adopted a view similar to Kant’s and introduces “positive” and “negative” just as relative terms, without ascribing one of them to expansion and the other to contraction. At this time, Ørsted did not accept Schelling’s ideas: The two attempts to build a chemistry on the basis of the critical metaphysics of nature that I am familiar with are so unsuccessful that they have brought their authors into the most evident contradiction with its foundations. The first to make an attempt of this kind is, as far as I know, the above-mentioned Eschenmayer, who builds it on the doctrine of the relation between the fundamental forces of matter which we have seen above. In his Ideen zur Philosophie der Natur, Schelling has adopted the same doctrine and developed it more precisely. As the theory of the former philosopher is false from its first foundation, the chemistry which he has based upon it is also false and is in conflict with the basic ideas of dynamics. As Schelling tries to develop the same chemical theory on other grounds, I only want to demonstrate its incorrecteness by means of a few observations.173

Towards the end of this work, Ørsted mentioned two of Schelling’s books: Ideen zu einer Philosophie der Natur, and Von der Weltseele, and remarked that “these two books certainly deserve attention because of the beautiful and grand ideas which are found in them, but the insufficiently rigorous method, whereby the author adds empirical theorems without distinguishing them adequately from a priori theorems, deprives the book of much of its value, in particular because the empirical theorems that he adduces are often completely false.”174 At this time, it was impossible to classify Ørsted as a follower of Schelling’s ideas. Shortly afterwards, however, Ørsted’s opinion about Schelling began to change. In the same year (1799) he published his “Dissertation on the structure of the elementary metaphysics of external nature” where he presented a favourable attitude: This essay of mine was almost finished when Schelling’s excellent Erster Entwurf einer Naturphilosophie arrived here, so I could not use it in this place, which I certainly regret; in any case, his book contributes much more to the higher than to the elementary metaphysics of nature. What I have tried to establish in this dissertation about the force of cohesion is in accordance with the views of this philosopher; I have not, however, derived these findings from his book […]175

Although Ørsted highly praised Kant, at some places he openly criticised him: “[…] although I originally intended to follow in Kant’s footsteps as far as this subject is concerned, when I thought it over more carefully I was forced to leave that trail.”176 Schelling’s influence upon Ørsted became stronger after 1802. In 1802, Ørsted’s ideas were already regarded as related to Naturphilosophie, and this was a cause of concern around him. In Berlin, during his continental travels (1801–1802), 172

173 174 175

176

Ørsted, “Fundamentals of the metaphysics of nature,” (1799), in Selected scientific works…, §39, p. 61. Ibid. §67, p. 71. Ibid. §80, p. 77. Ørsted, “Dissertation on the structure of the elementary metaphysics of external nature,” (1799), in Selected scientific works…, pp. 79–80. Ibid. p. 84.

378

R. DE A. MARTINS

he defended Naturphilosophie against the criticisms of Alexander Nicolaus von Scherer.177 One decade later, Ørsted contrasted Kant’s and Schelling’s contributions to physics in a very suggestive way: The progress of philosophy in the eighteenth century has not been without influence on general physics. The perspicacity of Immanuel Kant liberated it from the atomistic system, which, though of speculative origin, was made the basis of experimental physics. F. W. J. Schelling created a new natural philosophy, the study of which must be important to the empirical student of nature and must both inspire many new ideas in him and also prompt him to re-examination of much that was previously considered unquestionable.178

Notice that Ørsted did not emphasise Kant’s contribution to the dynamical viewpoint, but only his anti-atomism. Notice also that his words present Schelling’s contribution as much more relevant than Kant’s. It is also relevant to point out that Ørsted ascribed to Schelling—not to Kant— the attempt to find the unity behind all phenomena: […] As none of the physical processes is completely isolated but is connected with others, it follows that the science which we are discussing here cannot be divided into two parts, like physics itself, but that it must constitute a single, organic science, in relation to which experimental physics only serves as a means. We have fragments of such a science, for example, physical astronomy, geology, and meteorology, but the complete science does not exist yet and can never be reached by the path of experience. It is well-known that Schelling, through speculation, has produced an attempt which, as such, is of incalculable value, but the combined efforts of a great number of blessed geniuses are probably required for the accomplishment of this task.179

In his writings, Ørsted not often refers to Schelling by name. However, there is a very strong influence that can be noticed when one compares the content of their ideas. Let us show just one instance: Ørsted’s description of magnetism and electricity as related to one and two dimensions: A brief outline of what we know about the effects of these forces is sufficient to show us the possibility that all the different forces of nature can be traced back to those two fundamental forces. How could there be three more different effects than heat, electricity and magnetism! Yet, all of these are due to the effect of the same fundamental forces, only in different forms. Magnetism acts only in a line which is determined by the two opposite poles and the intermediate point of equilibrium. Purely electrical effects only follow surfaces. Heat works equally freely in all directions in a body.180

It is possible to find very similar ideas in Schelling: “[…]magnetism, as a process, as form of activity, is the process of length, electricity the process of breadth, just as the chemical process, on the other hand, is that which alone affects cohesion or form in all dimensions, and hence in the third.”181 Or, more fully:

177 178 179

180 181

Jacobsen, Between Naturphilosophie and Tradition…, pp. 18, 40–42. Ørsted, “First introduction to general physics,” (1811), in Selected scientific works…, p. 305. Ørsted, “New investigations into the question: What is chemistry?,” (1805), in Selected scientific works.., p. 199. Ibid. p. 197. Schelling, Ideas for a philosophy of nature, …, p. 137.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

379

What was cohesion and magnetism in the first and second potency, returns here, after the ideal principle has identified itself with matter for the first dimension, as the formative impulse, as reproduction. What there presented itself as relative cohesion, or electricity, is here, in the absolute identification of form and matter for the second dimension, raised to irritability, to the living power of contraction. Finally, where the light takes the place of matter altogether, and presses into the third dimension, so that essence and form in this way become wholly one, the chemical process of the lower potency passes over into sensibility, into the inner absolute formative power.182

Taking into account all evidence presented here, it seems that Ørsted was strongly influenced by Schelling’s Naturphilosophie during the first decade of the 19th century. At times this influence was direct. More often, however, he was influenced through Ritter’s work.

18. CONCLUDING REMARKS In the early decades of the 19th century no magnetochemical effect had become reproducible—or, in Ian Hacking’s terminology, no magnetochemical phenomenon had been created: To experiment is to create, produce, refine and stabilize phenomena. If phenomena were plentiful in nature, summer blackberries there just for the picking, it would be remarkable if experiments didn’t work. But phenomena are hard to produce in any stable way. That is why I spoke of creating and not merely discovering phenomena. That is a long hard task.183

During that period, the search for chemical effects of magnetism was driven by two different impulses. The first influence, that acted upon Ritter (and Arnim), was the belief in a fundamental unity of all forces of nature and the search for definite relationships between the polarities of those forces. Ritter’s magnetochemical investigations can only be fully understood in the philosophical context of Schelling’s Naturphilosophie, that guided his experimental research. By Ritter’s personal influence, and due to his sharing the main Romantic tenets, Ørsted came to accept all the effects he described as genuine, and helped to disseminate Ritter’s discoveries. Ritter’s magnetochemical researches were criticised by Paul Erman in 1807, as described above. It is noteworthy, however, that Erman’s attack was not an isolated and neutral piece of scientific work. Erman’s papers were published in the Annalen der Physik, where there appeared, at the same time, severe attacks against Naturphilosophie. The speculative method defended by Schelling was condemned by Ludwig Wilhelm Gilbert, the editor of the Annalen der Physik. Gilbert asserted that the vogue of galvanism had passed. He strongly criticised the abuse of “duality” and “polarity” in all fields of chemistry and physics. Erman added that Schelling’s Naturphilosophie was a greater blame to the Germans than twenty defeats by Napoleon.184 182 183

184

Ibid. p. 138. Ian Hacking, Representing and Intervening. Introductory topics in the philosophy of natural science (Cambridge: Cambridge University Press, 1983), p. 230. Stuart Strickland, “Galvanic disciplines: the boundaries, objects, and identities of experimental science in the era of Romanticism,” History of Science 33 (1995), pp. 449–468, at 452; Armin Hermann, “Unity and metamorphosis of forces …, p. 56.

380

R. DE A. MARTINS

Notice that Gilbert’s journal had published many papers of the Romantic physicists—including Ritter. It seems that Ritter’s speculations about the divining rod triggered Gilbert’s criticism against this approach in 1807.185 The attack against Ritter and Naturphilosophie in 1807 by Gilbert and Erman was successful, and for many years no new attempt was made to find a relation between magnetism, galvanism, and chemical phenomena. Ørsted was probably one of the very few people who in the 1810s still entertained expectations concerning the unity of all forces of nature. Maschmann’s work was not inspired by philosophical beliefs. It was due to an accidental observation. It seems that Maschmann, Hansteen, and Ørsted felt insecure about the reality of these effects before 1820, since they did not publish any account of those experiments. Perhaps the criticism suffered by Ritter one decade earlier had some bearing on this cautious silence. After 1820 the situation changed, and many researchers turned to magnetochemical experiments, not as the result of new philosophical influences, but as an effect of the unexpected discovery of electromagnetism. The situation was similar to what happened from 1896 onwards, after the discovery of X-rays. I agree with Oliver Lodge, who remarked that new discoveries usually produce general doubts about accepted knowledge, and speculative activity.186 Up to 1812 Ørsted had a firm assurance that Ritter’s experiments had demonstrated the relation between galvanism and magnetism. As remarked by Anja Jacobsen, the very name of the book he published at this time (in French: Recherches sur l’identité des forces chimiques et électriques) shows that he still accepted one of the central ideas of Schelling’s Naturphilosophie at this time.187 It is difficult to ascertain whether Ørsted’s later denial of Ritter’s findings was due to his most intimate conviction that Ritter was a poor experimenter, or a response to changing cultural forces. In 1830 he was content to accept Maschmann’s and Hansteen’s experiments, although the influence of magnetism upon the formation of Diana’s tree was controversial. This paper did not directly address the general problem of Ørsted’s relation to Kant. It is possible that in his earliest and later periods Ørsted was more strongly associated to Kant’s ideas than to Schelling’s, as claimed by Dan Christensen.188 The contention of this paper is that during his early scientific career, in the course of publishing his accounts of Ritter’s experimental researches on the polarities of nature, Ørsted’s ideas had a clear Nature-philosophical inspiration. Ritter’s search for definite relations between the magnetic poles and the polarities of other natural forces cannot be understood apart from his fundamental philosophical beliefs. Ørsted’s presentation of Ritter’s ideas and experiments, together with his later favourable comments upon those researches, is a strong evidence that he was also guided by very similar ideas at that time. Ritter’s emblematic way of thinking,

185 186

187 188

Kaiser, “Symmetries in Romantic physics…, p. 86. Oliver Lodge, “The discovery of radioactivity, and its influence on the course of physical science (Becquerel memorial lecture),” Journal of the Chemical Society 101 (1912), 2005–2031. Jacobsen, Between Naturphilosophie and tradition…, p. 149. Dan Charly Christensen, “Ørsted’s concept of force,” (this volume).

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

381

shared by Ørsted, also points out a Romantic influence that cannot be ascribed to Kant. Altogether, this specific case study supports the contention of a strong influence of Schelling’s Naturphilosophie in Ørsted’s early scientific career. State University of Campinas, Sao Paolo, Brazil

ACKNOWLEDGMENTS The author is grateful to the Brazilian National Council for Scientific and Technological Development (CNPq) and to the São Paulo Research Foundation (FAPESP) for supporting this research.

BIBLIOGRAPHY [Anonymous]. “Extrait d’une lettre écrite de Munich au Prof. Pictet sur quelques expériences galvanicomagnétiques, faites récémment par Mr. Ritter.” Bibliothèque Britannique, ou Recueil. Sciences et Arts 31: 97–100, 1806.189 [Anonymous]. “On the influence of magnetism on chemical action.” In Encyclopaedia Britannica. 8th. ed. vol. 14, pp. 41–42 Edinburgh: Adam & Charles Black, 1857. Aesch, Alexander Gode-Von. Natural science in German romanticism. New York: AMS Press, 1966. Arnim, Ludwig Achim von. “Ideen zu einer Theorie des Magneten.” Annalen der Physik 3: 48–64, 1800. —— . “Bemerkungen über Volta’s Säule.” Annalen der Physik 8: 163–196, 257–284, 1801. Ashworth, Jr., William B. “Natural history and the emblematic world view.” In Lindberg, David C. and Westman, Robert S. (eds.), Reappraisals of the Scientific Revolution. Cambridge: Cambridge University Press, pp. 303–332, 1990. B. M. “Observations on Mr. Murray’s paper on the decomposition of metallic salts by the magnet.” The Annals of Philosophy [ser. 2] 3: 39–41, 1822 (a). —— . An answer to Mr. Murray’s “Reply”. The Annals of Philosophy [2] 3: 384–385, 1822b. Bequerel, Antoine César. Traité expérimental de l’électricité et du magnétisme et de leurs rapports avec les phénomènes naturels. 7 vols. Paris: Firmin Didot, 1834–1840. Bequerel, Antoine César and Bequerel, Edmond. Traité d’électricité et de magnétisme et des applications de ces sciences a la chimie, a la physiologie et aux arts. 3 vols. Paris: Firmin-Didot, 1855–1856. Bernoulli, Christian. New galvanic experiments by M. Ritter. Extracted from a letter from M. Christ. Bernoulli. Philosophical Magazine 23: 51–54, 1806.190 Berzelius, Jöns Jacob. Jahresbericht über die Fortschritte der physischen Wissenschaften. 30 vols. Tübingen: H. Laupp, 1822–1851. Bhatnagar, Shanti Swarupa and Mathur, K. N. Physical principles and applications of magnetochemistry. London: MacMillan, 1935. Caneva, Kenneth L. Ørsted’s presentation of others’—and his own—work. 2004 (this volume) Canallo, Tiberius. “Magnetical experiments and observations.” Philosophical Transactions of the Royal Society of London 76: 62–80, 1785; 77: 6–25, 1786. Christensen, Dan Ch. “The Ørsted-Ritter partnership and the birth of Romantic natural philosophy.” Annals of Science 52: 153–185, 1995. 189

190

This notice was reproduced in a number of journals: Aus dem Intelligenzblatte der Allgem. Litterat. Zeit. Den 5ten Febr. 1806, Annalen der Physik 22 (1806), pp. 223–224; “Extract of a letter to professor Pictet, from a Correspondent at Munich, upon some galvanico-magnetic experiments recently made by M. Ritter,” Philosophical Magazine 25 (1806), pp. 368–369, 1806; Estrato d’una lettera scritta da Monaco in Baviera al Sig. Prof. Picted di Ginevra su alcuni sperimenti galvanico-magentici fatti recentemente dal Sig. Prof. Ritter, Nuova Scelta d’Opuscoli 1 (1806), pp. 334–336. Cf. the Italian version of the letter see Bernouolli, Christian. ‘Scoperte galvaniche del Sig. Ritter estratte da una lettera del Sig. Cristoforo Bernoulli. Nuova Scelta d’Opusculi Interessanti sulle Scienze e sulle Arte 1: 201–202, 1806.

382

R. DE A. MARTINS

Christensen, Dan Charly. Ørsted’s concept of force. 2004 (this volume) Cruickshank, William. “Some experiments and observations on galvanic electricity.”191 A Journal of Natural Philosophy, Chemistry, and the Arts [series 1] 4: 187–191, 1801. Dulk, Friedrich Philipp. “Ueber die chemische Einwirkung des Magnetismus.” Archiv für die gesammte Naturlehre 6: 457–467, 1825. Erdmann, Otto Linné. “Versuche über den angeblichen Einfluss des Magnetismus auf chemische Wirkungen und auf den Krystallisationsprocess.” Journal für Chemie und Physik192 56: 24–53, 1829a. —— . “Expériences sur l’influence présumée du magnétisme sur les effets chimiques et la marche de la cristallisation.” Bibliothèque Universelle des Sciences, des Belles-Lettres et Arts. Sciences et Arts 42: 96–103, 1829b. —— . “On the supposed influence of magnetism in the phenomena of chemical combinations and crystallizations.”193 The American Journal of Science and Arts 18: 395–397, 1830. Erman, Paul. “Beiträge über electrisch-geographische Polarität, permanente electrische Ladung, und magnetisch-chemische Wirkungen.” Annalen der Physik 26: 1–35, 121–145, 1807. Fresnel, Augustin. “Note sur des essais ayant pour but de décomposer l’eau avec un aimant.” Annales de Chimie et de Physique [series 3] 15: 219–222, 1820. Gmelin, Leopold. Handbook of chemistry. Transl. Henry Watts. 18 vols. plus index. London: Cavendish Society, 1848–1872. Granet, Marcel. La Pensée Chinoise. Paris: La Renaissance du Livre, 1934. Hachette, Jean Nicolas Pierre. “Expérience sur le magnétisme de la pile électrique.” Correspondance sur l’École Impériale Polytechnique, a l’usage des élèves de cette École, 1: 151–153, 1805. —— . “Sur les expériences électro-magnétiques de MM. Ørsted et Ampère.” Journal de Physique, de Chimie et d’Histoire Naturelle 91: 161–166, 1820.194 Hacking, Ian. Representing and intervening. Introductory topics in the philosophy of natural science. Cambridge: Cambridge University Press, 1983. Hansteen, Christopher. “Wiederholung und Bestätigung der Versuche durch Hrn. Prof. Hansteen.” Annalen der Physik 70: 239–242, 1822. —— . “Répétition et confirmation des mêmes expériences, par le professeur Hansteen.” Annales de Chimie et de Physique [series 3] 38: 206–208, 1828. Hermann, Armin. “Unity and metamorphosis of forces (1800–1850): Schelling, Ørsted, and Faraday.” In Doncel, Manuel G., Hermann, Armin, Michel, Louis and Pais, Abraham (eds.). Symmetries in Physics (1600–1980). Barcelona: Universitat Autònoma de Barcelona, pp. 51–62 1987. Hesse, Mary. “Models and analogy in science.” In Edwards, Paul (ed.). The Encyclopedia of Philosophy. New York: Macmillan & The Free Press, vol. 5, pp. 354–359, 1967. Humboldt, Frederick Alexander von. Versuche über die gereizte Muskel- und Nervenfaser, nebst Vermuthungen über den chemischen Process des Lebens in der Thier- und Pflanzenwelt. Posen: Decker und Compagnie, 1797. 2 vols. —— . Expériences sur le galvanisme et en général sur l’irritation des fibres musculaires et nerveuses. Traduction de l’allemand [by Gruvel or Gravel], publiée, avec des additions, par J. Fr. N. Jadelot. Paris: De l’imprimerie de Didot jeune, an VII, 1799. Jacobsen, Anja Skaar. Between Naturphilosophie and tradition: Hans Christian Ørsted’s dynamical chemistry.195 Aarhus: University of Aarhus, 2000. Jacyna, L. S. “Galvanic influences: themes in the early history of British animal electricity.” In Bresadola, Marco and Pancaldi, Giuliano (eds.). Luigi Galvani International Workshop. Proceedings. Bologna: Università di Bologna, pp. 167–185 1999. Kaiser, Walter. “Symmetries in Romantic physics.” In Doncel, Manuel G., Hermann, Armin, Michel, Louis and Pais, Abraham (eds.). Symmetries in Physics (1600–1980). Barcelona: Universitat Autònoma de Barcelona, pp. 77–92 1987. Kastner, Karl Wilhelm Gottlob. “Zur Geschichte des Galvanismus.” Archiv für die gesammte Naturlehre 6: 442–452, 1825. 191

192

193 194

195

This paper was translated into German as: Versuche und Beobachtungen über einige chemische Wirkungen der galvanischen Electricität. Annalen der Physik 6: 360–8, 1800. This periodical had two different titles. It was also called Jahrbuch der Chemie und Physik, from vol. 30 onwards (with a new numbering of the volumes). Journal für Chemie und Physik 56 corresponded to Jahrbuch der Chemie und Physik 26. This is a short account of Erdman’s paper in the Bibliothèque Universelle. There is an English translation of this paper: On the electromagnetic experiments of MM. Ørsted and Ampere. By Mr. Hatchett [sic]. The Philosophical Magazine and Journal 57: 40–44, 1821. Unpublished Ph.D. thesis.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

383

—— . “Chemische Gegenwirkung des magnetischen Eisens.” Nachtrag zum Vorhergehenden. Archiv für die gesammte Naturlehre 15: 336–344, 1828. Kirwan, Richard. “Thoughts on magnetism.” A Journal of Natural Philosophy, Chemistry, and the Arts [1] 4: 90–94, 133–135, 1801. Lee, Everett S. and Poust, C. M. “Measurement of surge voltages on transmission lines due to lightning.” General Electric Review 30: 135–145, 1927. Lehot, C. J. [M. Lehot adresse une lettre relative à l’expérience dont M. Fresnel a rendu compte à l’Académie dans la dernière séance]. Annales de Chimie et de Physique [series 3] 15: 406–408, 1820. Lichtenberg, Georg Christoph. De novo methodo naturam ac motum fluidi electrici investigandi. Novi Commentarii Societatis Regiae Scientiarum Gottingensis. Commentationes physicae et mathematicae classis, 8: 168–180, 1777. Lloyd, G. E. R. “Analogy in early Greek thought.” In Wiener, Philip P. (ed.). Dictionary of the history of ideas: studies of selected pivotal ideas. New York: Charles Scribner’s Sons, vol. 1, pp. 60–63, 1973. Lodge, Sir Oliver. “The discovery of radioactivity, and its influence on the course of physical science” (Becquerel memorial lecture). Journal of the Chemical Society, 101: 2005–2031, 1912. Lüdicke, August Friedrich. “Versuche mit einer magnetischen Batterie.” Annalen der Physik 9: 375–378, 1801. —— . “Fortsetzung der Versuche mit verbundnen Magnetstählen, und ein paar Bemerkungen zu Volta’s Säule.” Annalen der Physik 11: 114–119, 1802. —— . “Ueber den Einfluss des magnetischen Wirkungskreises auf die Krystallisation einiger Salze.” Annalen der Physik 68: 76–83, 1821. Martins, Roberto de Andrade. “Jevons e o papel da analogia na arte da descoberta experimental: o caso da descoberta dos raios X e sua investigação pré-teórica.” Episteme. Filosofia e História das Ciências em Revista 3 (6): 222–249, 1998. —— . “Romagnosi and Volta’s pile: Early difficulties in the interpretation of voltaic electricity.” In Bevilacqua, Fabio & Fregonese, Lucio (eds.). Nuova Voltiana: Studies on Volta and his Times. Pavia/Milano: Università degli Studi di Pavia / Ulrico Hoepli, vol. 3, pp. 81–102, 2001. Maschmann, Hans Henrik. “Einwirkung des Erdmagnetismus auf Auscheidung des Silbers.” Annalen der Physik 70: 234–239, 1822. —— . “Effets du magnétisme terrestre sur la précipitation de l’argent, observés par M. Muschman” [sic]. Annales de Chimie et Physique [series 3] 38: 201–205, 1828a. —— . “Effect of magnetism on the precipitation of silver.” Quarterly Journal of Science 26: 429–431, 1828b. Meyer, Kirstine. “The scientific life and works of H. C. Ørsted.” In Kirstine Meyer (ed.) Ørsted, H. C. Scientific Papers. Copenhagen, Andr. Fred. Høst & Søn, vol. 1, 1920. Mottelay, Paul Fleury. Bibliographical History of Electricity and Magnetism. London: Charles Griffin, 1922. Murray, John. “On the decomposition of metallic salts by the magnet.” London, Edinburgh and Dublin Philosophical Magazine 58: 380–382, 1821. —— . Reply to B. M. The Annals of Philosophy [ser. 2] 3: 121–123, 1822. Nicholson, William. Scientific news. A Journal of Natural Philosophy, Chemistry, and the Arts [series 2] 1: 234–236, 1802. Ørsted, Hans Christian. “Expériences sur un appareil à charger d’électricité par la colonne éléctrique de Volta; par M. Ritter, à Jena. Présentées à l’Institut National par J. C. Ørsted, docteur à l’Université de Copenhague.” Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts 57: 345–368, 1803a (an XII de la République). —— . “Expériences avec la pile électrique faites para M. Ritter, à Jéna; communiquées par M. Ørsted.” Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts 57: 401–405, 1803b (an XII de la République).196 —— . “Expériences sur le magnétisme, par M. Riter, à Jéna, communiquées par Ørsted, docteur à l’université de Copenhagen.” Journal de Physique, de Chimie, d’Histoire Naturelle et des Arts 57: 406–409, 1803c (an XII de la République).197

196

197

A translation was published in Nicholson’s Journal: Experiments with the electric pile, by Mr. Ritter, of Jena. Communicated by Mr. Ørsted. A Journal of Natural Philosophy, Chemistry, and the Arts [2] 8: 176–180, 1804. A translation was published in Nicholson’s Journal: Experiments on magnetism; by Mr. Ritter, of Jena. Communicated by Dr. Ørsted, of Copenhagen. A Journal of Natural Philosophy, Chemistry, and the Arts [2] 8: 184–186, 1804.

384

R. DE A. MARTINS

—— . “Ueber der neuesten Fortschritte der Physik. Europa. Eine Zeitschrift.” 1 (2): 20 ff., 1803 (d). Reproduced in Kirstine Meyer (ed.) Ørsted, H. C. Scientific Papers. Copenhagen, Andr. Fred. Høst & Søn, vol. 1, pp. 112–131 1920. —— . Ansicht der chemischen Naturgesetze durch die neueren Entdeckungen gewonnen. Berlin: Realschulbuchhandlung, 1812. Reproduced in Kirstine Meyer (ed.) Ørsted, C. H. Naturvidenskabelige Skrifter. København: Andr. Fred. Høst & Søn, vol. 2, pp. 35–169 1920. —— . Experimenta circa effectum conflictus electrici in acum magneticam. Hafniae: Schultz, 1820.198 —— . “Expériences de Ritter, analysées par M. Ørsted.” Annales de Chimie et de Physique [series 3] 38: 197–200, 1828. —— . “Thermo-electricity.” In Brewster, D. (ed.). The Edinburgh Encyclopaedia. 18 vols. Edinburgh, vol. 18, pp. 573–589 1830. —— . Selected scientific works. Translated and edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen. Introduction by Andrew D. Wilson. Princeton, NJ: Princeton University Press, 1998. Oken, Lorentz. Lehrbuch der Naturphilosophie. Jena: Friedrich Frommann, 1831. Ostwald, Wilhelm. Elektrochemie. Ihre Geschichte und Lehre. Leipzig: Von Veit, 1896. Pera, Marcello. The ambiguous frog: the Galvani-Volta controversy on animal electricity. Trad. J. Mandelbaum. Princeton, NJ: Princeton University Press, 1992. Rendu, Louis. “Observations qui tendent a prouver que la cristallisation de tous le corps est un phénomène électrique.” Bibliothèque Universelle des Sciences, Belles Lettres et Arts. (Sciences et Arts) 38: 304–317; 39: 58–72, 1928a. —— . “Influence du magnétisme sur les actions chimiques.” Annales de Chimie et de Physique [series 3] 38: 196–197, 1828b. Ridolfi, C. Lettera all’editore dell’Antologia. Antologia 7: 498–501, 1822. Ritter, Johann Wilhelm. “Versuche und Bermerkungen über den Galvanismus der Voltaischen Batterie.” Annalen der Physik 7: 431–484, 1801. —— . “Auszüge aus Briefen verschiednen Inhalts an der Herausgeber. 1. Von Herrn J. W. Ritter.” Annalen der Physik 15: 106–110, 1803a. —— . “Einiges über Nordlichter und deren Periode, und über den Zusammenhang des Nordlichts mit dem Magnetimus, und des Magnetimus mit den Feuerkugeln, dem Blitze und der Electricität.” Annalen der Physik 15: 206–226, 1803b. —— . Das electrische System der Körper. Leipzig: C. H. Reclam, 1805. —— . Fragmente aus dem Nachlass eines jungen Physikers. Leipzig: Infel Verlag, 1938. Ruhland, Reinhold Ludwig. “Vermischte Bemerkungen electrischen und magnetischen Inhalts.” Journal für Chemie und Physik 11: 16–25, 1814. Schelling, Friedrich Wilhelm Joseph von. Ideas for a philosophy of nature. Translated by Errol E. Harris and Peter Heath. Cambridge: Cambridge University Press, 1995. Schweigger, Johann Salomo Christoph. “Ueber Cohäsion, in Abhängigkeit von krystall-elektrischer Anziehung.” Journal für Chemie und Physik 44: 79–86, 1825. Selwood, Pierce Wilson. Magnetochemistry. New York: Interscience, 1943; 2nd. ed. New York: Interscience, 1956. Shanahan, Timothy. Kant, “Naturphilosophie, and Ørsted’s discovery of electromagnetism: a reassessment.” Studies in the History and Philosophy of Science 20: 287–305, 1989. Snelders, H. A. M. “The influence of the dualistic system of Jakob Joseph Winterl (1732–1809) on the German Romantic era.” Isis 61: 231–239, 1970. Stauffer, Robert C. “Speculation and experiment in the background of Ørsted’s discovery of electromagnetism.” Isis 48: 33–50, 1957. Strickland, Stuart. “Galvanic disciplines: the boundaries, objects, and identities of experimental science in the era of Romanticism.” History of Science 33: 449–468, 1995. Sue, Pierre. Histoire du galvanisme et analyse des différens ouvrages publiés sur cette découverte, depuis son origine jusqu’à ce jour. 2 vols. Paris: Bernard, an X, 1802. Van Swinden, J. H. Analogie de l’éléctricité et du magnétisme ou récueil de mémoires, couronnés par l’Académie de Bavière. 3 vols., La Haye/Paris: La Compagnie/Veuve duch*esne, 1785. Volta, Alessandro. “On the electricity excited by the mere contact of conducting substances of different kinds.” Philosophical Transactions of the Royal Society 90: 403–431, 1800. Wedekind, Edgar. Magnetochemie. Berlin: Gebrüder Borntraeger, 1911.

198

Ørsted’s paper was reproduced in the Journal für Chemie und Physik 29: 275–281, 1820. It was reprinted in Larsen, A. (ed.). The discovery of electromagnetism made in the year 1820 by H. C. Ørsted. Copenhagen, 1920. This book also contains facsimile reprints of early translations (German, French, English, Italian) of Ørsted’s work.

ØRSTED, RITTER, AND MAGNETOCHEMISTRY

385

Wetlzar, Gustav. “Ueber den elektrodynamischen Zustand, welchen Eisen und Stahl durch Berührung mit saurer salpetersaurer Silberslösung oder reiner Ammoniakflüssigkeit erlangen.” Journal für Chemie und Physik 56: 206–227, 1829. Wiedeman, Gustav. Die Lehre von der Elektricität. 3 vols. Braunschweig: Friedrich Vieweg und Sohn, 1882–1883. Wilson, Andrew D. Introduction. In Ørsted, Hans Christian. Selected scientific works. pp. xv–xl Princeton, NJ: Princeton University Press, 1998. Yelin, Ritter von. “Ueber den Zusammenhang der Electricität und des Magnetismus.” Annalen der Physik 66: 395–411, 1820. Zantedeschi, Francesco. “Nota sopra l’azione della calamita e di alcuni fenomeni chimici.”199 Biblioteca Italiana o sia Giornale di Letteratura, Scienze ed Arti 53: 398–402, 1829. —— . Relazione delle principali scoperte magneto elettriche (Opuscoli Fisici di Varii Autori, vol. 1, n. 36), Verona: Antonelli, 1834.

199

Cf. the French translation of this article: Note sur l’influence mutuelle du magnétisme et des actions chimiques. Bibliothèque Universelle des Sciences, Belles Lettres et Arts 1: 22–29, 1830.

ØRSTED’S WORK ON THE COMPRESSIBILITY OF LIQUIDS AND GASES, AND HIS DYNAMIC THEORY OF MATTER OLE KNUDSEN

In this paper I shall attempt to answer the question why Ørsted should have devoted so much time and effort to measuring the compressibility of liquids and gases. In fact, in Ørsted’s scientific oeuvre his experimental researches on the compressibility of water and gases stand out as the most long-lasting, sustained, and painstaking research effort of all. It occupied him off and on from 1818 to 1845, and for a period in the 1820s, just when one might have expected him to concentrate on exploring, like many of his colleagues around Europe were doing, the new field opened up by his discovery of electromagnetism, it seems to have taken up most of the time he could spare from his heavy teaching and administrative duties. What was it about this subject that so attracted Ørsted? I will argue that a partial answer may lie in Ørsted’s firm belief in Kant’s dynamic theory of matter and his opposition to the atomistic views that dominated particularly French physics and chemistry.1 As will appear in the following, there is very little direct evidence that his work on compressibility was related to his dynamic views; in his papers on compressibility Ørsted never made reference to the dynamic theory and in his discussions of the constitution of matter he mentioned the phenomenon of compressibility and elasticity of material bodies as an argument for the dynamic theory only rarely. I shall therefore restrict myself to giving an account of Ørsted’s work on compressibility and, subsequently, of the development of his views on the constitution of matter. In the final section I shall then attempt to evaluate the evidence, direct or indirect, for claiming that Ørsted may have seen his experimental results on compressibility as arguments for the dynamic view of matter against the atomistic.

1

For similar suggestions, see Anja Skaar Jacobsen: Between Naturphilosophie and Tradition: Hans Christian Ørsted’s Dynamical Chemistry, unpublished Ph.D. thesis, University of Aarhus 2000, pp. 27–28; and Ole Norling-Christensen: “Hans Christian Ørsted,” in: Dansk Biografisk Leksikon, 3rd ed., vol. 16. Copenhagen 1984, pp. 196–202, on p. 199.

387 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 387–398. © 2007 Springer.

388

O. KNUDSEN 1. ØRSTED ON COMPRESSIBILITY

Ørsted’s earliest remark on this subject is found in his review of the new edition of A. W. Hauch’s textbook.2 To Hauch’s claim that Abich and Zimmermann (of whom more below) had proved that water possesses elasticity3 Ørsted objected that the diminution of volume in their experiments might be due to “expulsion of the matter of heat by compression” and that Abich and Zimmermann who “have met so many other objections have not eliminated this one.”4 However, it took almost two decades before he took up the subject in earnest; in the meantime he had familiarized himself with the earlier work of Canton as well as that of Abich and Zimmermann. John Canton was the first to prove, in 1761, that water is compressible at all. He placed the water in a bottle with a long narrow neck and saw that the water level rose in the neck when the bottle was placed under the receiver of an air pump. When the pressure fell from one atmosphere to zero, the water expanded by a relative amount of 45 millionths, and when he increased the pressure from one to two atmospheres the water contracted by a similar amount. In 1779 a professor in Brunswick, E. A. W. Zimmermann, published a book on the elasticity of water in which he first gave a critical review of a large number of previous, mostly unconclusive, publications on this subject, including also a detailed account of Canton’s experiments.5 The last third of the book was devoted to a description of a “machine” which R. A. Abich, Obersalzinspector in Brunswick, had constructed after some unsuccessful attempts, and to an account of his and Zimmermann’s joint experiments on the compressibility of boiled and unboiled well water, saturated salt water, milk, and spirit of wine. As shown in Fig. 1, their apparatus consisted of a solid brass cylinder fitted with a piston T that could move up and down within an iron support fixed to the cylinder by screws. At first they forced the piston down by turning a screw (see Fig. 1); later, in order to measure the force on the piston they built the cylinder into a wall (Fig. 3) and used a weight, P, on a one-armed lever with fulcrum O also fixed in the wall. From the weight and position of the centre of gravity of the lever and the known weight P, which was increased in steps of a quarter of a hundredweight from 0 to 1¾ cwt., Zimmermann could calculate, using the law of the lever, that the force on the piston increased from 745 to 4862 Brunswick pounds. He did not give the corresponding pressures in atmospheres, but he did calculate that the weight of the atmosphere was equivalent to a weight of 8 pounds acting on the piston and this allows us to say that the pressures in the experiment ranged from 93 to 608 atmospheres. He gave the corresponding depressions of the piston in 144th parts of a Brunswick inch as ranging from 39 to 170 (or from

2

3 4 5

A. W. Hauch: Begyndelses-Grunde til Naturlæren [Foundations of Physics], anden og forbedrede Udgave [2nd and improved ed.], I–II, Copenhagen 1798–1799. Op. cit. II, p. 488. KM III, p. 36. E. A. W. Zimmermann: Ueber die Elasticität des Wassers, theoretisch und historisch entworfen. Zugleich eine Ankündigung eine hiehergehörigen Maschine. Leipzig 1779.

ØRSTED’S WORK ON THE COMPRESSIBILITY

389

Fig. 1. Abich’s compression apparatus. From Zimmermann, op. cit., n.6, plate II. 0.27” to 1.18”) for the experiments on unboiled well water; he also stated that the decreases in the volume of the water ranged from 0.1875 to 1 cubic inches.6 Zimmermann’s values give relative compressions of between 6.1 × 10−5 and 8.9 × 10−5 per atmosphere. The mean between these extremes, 7.5 × 10−5, was quoted 6

He found these values by measuring the amount of water displaced by the piston when pressed down into a full container to depths equal to the above-mentioned depressions. An easier way would be to multiply the depressions by the area of the piston. From the stated diameter of the piston (0.764”) I have tried to calculate the changes in volume and find much smaller values than Zimmermann’s measurements.

390

O. KNUDSEN

in Gehler’s dictionary with the remark that this value was only of historical interest because of the several defects of Abich and Zimmermann’s experiments and results.7 Ørsted’s first report on his own investigations on the compressibility of water was published in 1818 in a short letter to J. S. C. Schweigger, inserted in the latter’s Journal für Chemie und Physik, and in a slightly extended version communicated to the Royal Danish Academy of Sciences.8 He described an apparatus he had had constructed, no doubt inspired by Abich’s design. It consisted of a wide, very thick brass cylinder ending in a narrow tube inside which a small piston could move, enabling him to reach a high pressure with a small force. His pressure gauge consisted of a strong, narrow glass tube, filled with air and screwed into a hole in the water cylinder. Using Boyle’s law, he could get the pressure by reading off the volume of the air in the tube. He found that the relative compression of the water followed Boyle’s law and amounted to 12 × 10−5 per atmosphere at a temperature of 12° C, or almost three times Canton’s value. He criticised Canton for not taking correct account of “the influence of heat” and claimed that if this were done Canton’s experiments would lead to a higher compressibility. He spoke with approval of Abich’s experiments but blamed Zimmermann for making elementary errors in his calculations; when these were corrected, he said, the results, like his own, turned out to agree with Boyle’s law. Both Abich’s and Ørsted’s experiments suffered from the deficiency that their brass cylinders were subjected to high pressure only on the inside and therefore might expand as a result of the pressure. This would lead to too high a value for the compression of the water. Canton’s bottle, on the other hand, was subjected to the same pressure on the outside as well as on the inside and therefore, as he explicitly remarked, “not altered in its dimensions.”9 In 1820 an American copper engraver, Jacob Perkins, described a piezometer (a name he was the first to use) of his own invention which used Canton’s principle of equal pressure on the inside and outside (Fig. 2).10 The piezometer itself is shown on the left; it had a plunger, D, with a loosely fitting ring, Ea. He filled it with water and put it inside a cannon, placed vertically and likewise filled with water. A pump was screwed into the top of the cannon and the pressure was measured by the weight necessary to keep the valve C shut. After submitting the piezometer to a pressure of 100 atmospheres he found, on removing it from the cannon, that the ring Ea had moved eight inches up on the plunger, indicating that the latter had compressed the water in the piezometer by about 1% when the pressure was highest. This was done in America just before Perkins set out for England; during his passage he took the opportunity to sink his piezometer into the sea to a depth of 500 fathoms (915 m) where the pressure again was about 100 atmospheres, and again he found the ring to have moved eight inches up the plunger. After arriving 7

8 9

10

Johann Samuel Traugott Gehler’s Physikalisches Wörterbuch, neu bearbeitet von Brandes. Gmelin. Horner. Muncke. Pfaff., vol. 3, Leipzig 1827, pp. 208–209. Ørsted, Works, p. 407, and pp. 409–410. John Canton: “Experiments to prove that Water is not incompressible”, Philosophical Transactions 52 (1762) 640–643, on p. 642. Jacob Perkins: “On the compressibility of Water,” Philosophical Transactions (1820) 324–330.

ØRSTED’S WORK ON THE COMPRESSIBILITY

391

Fig. 2. Perkins’s piezometers. From Annales de chimie et de physique 16 (1821) 327.

in England he constructed a new piezometer (Fig. 2, right) which he subjected to a pressure of about 326 atm under a hydraulic press and found that an additional 3½% of water had been forced into it through the valve on top. All of these experiments gave values for the relative compressibility of about 0.0001 per atmosphere or a little more, thus agreeing better with Ørsted’s result than with Canton’s.

392

O. KNUDSEN

In 1821 Ørsted again took up the subject and designed a new piezometer which combined the advantages of Canton’s and Perkins’s principle of equal pressures on both sides, Canton’s bottle with the narrow neck, and Perkins’s use of water instead of air to transmit the pressure.11 The new apparatus (Fig. 3) consisted of a strong glass cylinder ABCD with a pump EFGH on top. The central part was the glass bottle ab with a capillary tube neck ending in a small funnel in which Ørsted put a small drop of mercury separating the water in the bottle from that in the rest of the cylinder. The pressure gauge consisted of an air-filled glass tube cd, open at the bottom. When the pump was worked the diminution of the volume of the water in the bottle was measured by the position of the mercury drop in the neck relative to the brass scale efgh, and the pressure by the compression of the air in the glass tube as given by the position of the water level relative to the same scale. Ørsted gave a very detailed description of the apparatus as well as of the experimental procedure. For instance he took great precaution in eliminating the effect of temperature changes on his results. He had found that an increase of 1°C made the water rise 27 lines (i.e. 12ths of an inch) in the capillary tube, corresponding to 108 divisions on his brass scale; thus he could detect a temperature change of 1/400 of a degree. To measure the compression corresponding to a given pressure he therefore took care to have the pressure equal to that of the surrounding air both before and after the measurement and check that the level in the tube was the same. He also placed a Breguet thermometer in the apparatus and noted no temperature change even when he increased the pressure to 5 atmospheres. As for the outcome of the experiment he just stated that the average result of a great number of trials, from 1/3 to 6 atmospheres, was a relative compression of 0.000045 per atmosphere, close to Canton’s value, and that the compression varied proportionally with the pressure. In the summer of 1824 Ørsted collaborated with a certain Captain v. Suensson of the Royal Engineers in a series of experiments to see if they could detect deviations from Boyle’s law for gases at high pressures. Up to 8 atmospheres they used a long mercury column in glass tubes cemented together; at higher pressures they used the reservoir of an air gun and a charging machine, both belonging to King Frederik VI who had most graciously let them be used in the interests of science. They first determined the volume of the reservoir and then calculated that it contained .891 g of air at one atmosphere. They then charged it with as much air as it would hold, weighed it in water and found the weight of the air it contained. By means of a weighted arm pushing against a valve (Fig. 4) they determined the pressure by measuring the weight necessary to just open the valve. Letting out the air a little at a time, each time repeating the measurements, they determined a table of corresponding values of pressure and density up to 60 atmospheres. The quotients of density into pressure proved to be nearly constant (within 6%); hence their experiment was an excellent verification of Boyle’s law within the above pressure range. They also experimented, using a different apparatus, with sulphurous acid gas 11

Ørsted’s report to the Danish Academy was published in German in Schweigger’s Journal in 1822 (Works, pp. 453–456) and in a slightly altered French version in Annales de physique et chimie in 1823 (Works, pp. 457–461).

ØRSTED’S WORK ON THE COMPRESSIBILITY

393

Fig. 3. Ørsted’s piezometer. (From Ørsted: Works, p. 460). (SO2) and found that it obeyed Boyle’s law up to a pressure of 2.3 atmospheres where its volume began to decrease faster than this law indicates and the gas began to liquefy. Ørsted’s report on this investigation was first published in Schweigger’s Journal in 1825 under the title “Experiments to prove that Mariotte’s law is applicable to all kinds of gases; and at all degrees of pressure under which the gases retain their aëriform state.”12 The following year he inserted it almost verbatim as part one of a paper to the Royal Danish Academy entitled “Contribution to the determination of the law of the compression of bodies,”13 the second 12 13

Ørsted: Works, pp. 481–491. Ørsted: Works, pp. 493–517

394

O. KNUDSEN

Fig. 4. The air gun reservoir used by Ørsted and v. Suensson. (From Ørsted: Works, p. 496). part of which contained a repetition of his earlier work on the compression of water. In a short section “General Remarks on Compression” at the end Ørsted drew the conclusion, very far-reaching in view of the limited number of substances he had experimented with: If we then imagine the same body in all three states, the gaseous, the liquid, and the solid, we see that the compression of all of them is proportional to the compressing forces, and that there is only a break in this regularity at the point of transition from one of these states to the other. It is true that we do not yet have any very certain experiments which show that liquids other than water follow this law of compression but it is so natural that there is no great risk in assuming it for all liquids…14

Measurements of compressibility continued to occupy Ørsted in the following years. In 1826 he reported measurements on the compressibility of mercury, his best results being 1.2 × 10−6 per atmosphere,15 in 1828 he compared the compressibility of water in a lead and a glass bottle and found a difference of less than 2 × 10−6 per atmosphere,16 and in a letter to William Whewell in 1833 he reported on a number of improvements of his piezometer which had resulted in a more exact value of 46.77 millonths per atmosphere for the compressibility of water at 3.75 °C where the density of water is at a maximum, a value less than 1% larger than the modern one.17 He now attempted to correct his measurements at other temperatures for the development of heat, using an estimated value of 1/40°C for the increase of 14 15 16 17

Ørsted: Works, p. 511. Ørsted: Works, pp. 521–525. Ørsted: Works, pp. 537–538. A modern textbook quotes the value 46.4 × 10−6 atm−1 for the compressibility of water. See Hugh D. Young and Roger A. Freedman: University Physics, with Modern Physics, 10th. ed., San Fransisco (Addison-Wesley) 2000, p. 342.

ØRSTED’S WORK ON THE COMPRESSIBILITY

395

temperature per atmosphere, corresponding to an expansion of the water of 1.79 millionths per atmosphere. He had now pushed the pressure up to 65 atmospheres and still found proportionality between compression and pressure.18 His last communication on the subject was a report to the Royal Danish Academy in 1845 on his collaboration with his former student, now inspector of roads, Ludvig Colding, to determine by direct measurement the increase of the temperature of water caused by pressure. Using a thermoelectric element they found a value of 1/49.2°C per atmosphere instead of Ørsted’s earlier estimate of 1/49.2 °C.19

2. ØRSTED ON THE CONSTITUTION OF MATTER Ørsted’s fight against atomism began with his exchange with the Lord Stewart Hauch in 1798–99. His opening shot was fired in an anonymous review (1798) of the first volume of Hauch’s textbook where he blamed Hauch for sticking “to the atomistic system according to which bodies are composed of small particles, distinguished from one another only in respect to their figure, so that the difference between bodies would come only from the difference of the pores.”20 He argued that in order to explain why bodies fill space the atomists had to postulate an expansive force and thus to invoke one of the main tenets of the dynamic system, and further that the difference between mechanical mixture and chemical combination was much easier to explain from the dynamic than from the atomistic point of view. In his Fundamentals of the Metaphysics of Nature (1799) Ørsted brought in the phenomena of compressibility and elasticity in his discussion of the atomistic system versus Kant’s dynamic view of matter. First, the two systems led to opposite conclusions on the limit of compressibility: …the atomistic system…assumes the basic particles or atoms to be completely incompressible and bodies which are composed of these only compressible in so far as they contain empty spaces. The dynamic system attributes to all bodies a fundamental force with which they fill space, but this has a degree above which one can find a greater one which can overcome it in so far as it is confined to a smaller space. Thus, impenetrability is invincible according to the former system but not according to the latter.

In other words, if matter consisted of hard, impenetrable atoms separated by pores of empty space there ought to be a limit to the compression of a body, attained when the atoms had been squeezed into contact with one another. If matter, on the other hand, was constituted as a state of equilibrium between primordial attractive and repulsive forces there could in principle be no limit to the amount of compression of a body if it was acted on by a sufficiently large external pressure. Elasticity was also a point in favour of the dynamic system, for this system had the advantage that it …explains phenomena for which the former [the atomistic] is not sufficient. Thus it is impossible to give a complete explanation of elasticity by means of the atomistic system. It is true that we can give a fairly good explanation of the kind of elasticity 18 19 20

Ørsted: Works, pp. 586–592. Ørsted: Works, pp. 607–609. KM III, p. 28.

396

O. KNUDSEN which is based on the attractive force using this system as it assumes such a force, but it cannot explain the kind of elasticity which is based solely on compression and expansion, for the expansion following compression can no more be due to the empty spaces than to the small incompressible particles, which have no force at all except to attract and to resist.21

Clearly, impenetrable atoms separated by empty spaces did not suffice to explain compression and expansion, even if they were assumed to exert gravitational attractions on one another. In the dynamic view, on the other hand, the extension of a body was determined by an equilibrium between Kant’s attractive and repulsive Grundkräfte, and an external compressive force might naturally shift the balance until a new state of equilibrium was reached at a smaller volume. To the second volume of his textbook, published later in 1799, Hauch added an appendix in which he discussed some of the points raised in the anonymous review and in Ørsted’s Fundamentals of the Metaphysics of Nature. To Ørsted’s argument that unless atoms were supposed to possess a repulsive force the attractive force of gravitation would compress them into a space of zero extension,22 Hauch answered that this would be impossible because The attractive or compressing force can, by its action on the parts of matter, effect only that these [parts] fill up space completely, when this has happened it will be impossible to fill it even further, and a force, no matter how great, even an infinitely great force, will not be able to introduce even the least further particle of matter into this space…23

For this reason Hauch rejected the dynamists’ repulsive force as an unproven and unnecessary hypothesis. It was easy for Ørsted, in his (no longer anonymous) review of Hauch’s vol. II, to counter this argument by asking for an explanation of how the atoms were able to fill up space so completely that no further compression was possible. He accused the atomistic system of explaining the finite extension of matter by postulating atoms which themselves possessed finite extension, thus providing an evidently incomplete explanation.24 Clearly, both Hauch and Ørsted understood atomism to postulate the existence of incompressible and impenetrable “billard ball” atoms of finite size and to assume that the compression of matter was limited by the total volume of its constituent atoms. But whereas Hauch found this postulate easy to accept and almost self-evident, Ørsted felt that the incompressibility of atoms called for further explanation which could only be given in terms of a repulsive force, hence that even the atomists would have to fall back on the primordial forces of the dynamic system. It is well known that Ørsted throughout his career stuck to the general dynamic philosophy of nature that he embraced in his youth, and that he remained antagonistic to atomism in its more naïve, literal form, even if he on occasion allowed himself to speculate on the possibility that bodies might have constituents 21 22 23 24

Ørsted: Works, pp. 74–75. Ørsted: Works, p. 61. Hauch, op. cit. n. 2, vol. II, p. 771. KM III, pp. 40–46.

ØRSTED’S WORK ON THE COMPRESSIBILITY

397

of a finite size. Thus in a discussion, in 1822, of Dalton’s chemical doctrine he spoke of Grunddeele (fundamental constituents) and said that Dalton calls these constituents atoms, whereby this doctrine gets an appearance of belonging exclusively to the atomistic philosophy of nature; but nothing hinders us from assuming, also in the dynamic system, a fundamental magnitude for every force-relation, so that the latter cannot express itself in a smaller space.25

Presumably, this passage means that Ørsted was willing to allow “atoms” in the form of small, dynamic, packages of force, but not Dalton’s ‘billard ball’ atoms; in any case it shows that his belief in the dynamic system was unshaken by Dalton’s doctrine, the basic features of which he had no problem in accepting or using.26 He elaborated on this view of atoms seven years later in an extensive correspondence on atomism and dynamism with C. S. Weiss, his longtime friend and fellow Kantian. His first letter to Weiss reflects his mature view that “force atoms,” which he again called Grundtheile, were not in conflict with a Kantian dynamic view of matter. He further explained that these constituents might be imagined to be in oscillatory or rotatory motion, and that their space-filling property was a combined effect of their innate forces and their oscillations, which latter he identified as the cause of heat.27 In the same letter he reiterated his argument from Metaphysical Foundations that an atom could only maintain its space against other atoms by an action (Thätigkeit), in which case …the presupposed atoms are no atoms anymore, they fill their space by an action (Thätigkeit) that can always be imagined to be overcome by a larger [one] and does not keep the parts together invincibly. Thus, if we should be brought to assume that the forces which constitute matter were crowded together in certain fixed points, or otherwise distributed in space in a particular way, yet we are far from being atomists. The essence of the atomistic system is to assume a dead, rigid, inactive (unthätiges) entity as the primordial basis (Urgrund) of the material world.28

3. CONCLUSION It is evident that Ørsted continued to imagine an atomist as someone who, like Hauch in 1799, would base science on inert and impenetrable particles of finite size, separated by pores of empty space. To this view he opposed a Kantian, dynamic ontology according to which matter was constituted by primordial attractive and repulsive forces. In this ontology the volume of a given material body was the result of a state of equilibrium between the primordial forces; if the body was subjected to an external pressure this equilibrium would be displaced towards a diminished volume.

25

26 27

28

H. C. Ørsted: Udsigt over Chemiens Fremskridt siden det attende Aarhundredes Begyndelse (Survey of the progress of chemistry since the beginning of the Eighteenth Century), Tidsskrift for Naturvidenskaberne 1 (1822) 1–55. KM III, pp. 301–329, on p. 314. Cf. Jacobsen, op. cit. n. 7911, pp. 211–222. Ørsted to Weiss, [30 January 1829], in: M. C. Harding, ed.: Correspondance de H. C. Örsted avec divers Savants, tome I, Copenhagen 1920, pp. 286–289. Ibid. pp. 282–283.

398

O. KNUDSEN

Given that this was Ørsted’s understanding of the basic difference between atomism and dynamism, it would not have been unnatural for him to think of his experiments on compressibility as furnishing a possible experimental test between the two fundamental conceptions of matter. If experiment had shown, or at least indicated, a limit to the compression of a given substance, this would have been a sign that the atoms had been squeezed into direct contact so that no further compression was possible, and atomism would have been provided with a forceful argument. On the other hand, if Ørsted’s results, as he claimed in 1826 and 1833, showed that all three states of matter obeyed Boyle’s law and that air and water followed this law without deviation up to pressures over 60 atmospheres this could hardly fail to have strengthened his belief in the dynamic view of matter. I suggest therefore that Ørsted’s absorption in the subject of compressibility may be partly explained by a desire to show that the dynamic view of matter could withstand the test of experimental measurements of the highest degree of precision attainable. At the same time I am aware that, apart from its possible philosophical importance, experimental science in itself has its own fascination that I find clearly evinced in the evident pleasure with which Ørsted recounted the successive improvements of his equipment and the difficulties he had overcome in the course of his investigations. University of Aarhus

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE FREDERICK GREGORY

1. This paper began as something different from what it has turned out to be. I began by thinking I would try to show what aspects of Schelling’s work were influential in shaping Ørsted’s understanding of science and religion, suspecting that it was Schelling and not Kant who was important here. I will do some of that; in fact, I now think both Kant and Schelling were significant influences, although in different ways. But what I found myself really doing as I proceeded was thinking about Ørsted’s uniqueness, and that led me to think more about ways that Ørsted reflected wider issues in the development of science during the first half of the 19th century. To begin with, Ørsted’s career spanned a most amazing period, from the beginning of the century to its midpoint. The changes in German science in this period, to say nothing of the changes in the German states themselves, were profound. There used to be a standard narrative that had German natural science flirt briefly with Naturphilosophie only to discard its temporary infatuation with it sometime after 1815 or so in favor of a more empirically based, experimental tradition that grew steadily over the remainder of the century. With the establishment of the journal Isis in 1817 and the founding of the Gesellschaft Deutscher Naturforscher und Ärzte in 1822, a German scientific community began to congeal. After 1848 the status of natural science had risen to such a degree that its presence in society was clearly visible, not only in the increasing wonders associated with new phenomena, like the amazing expectations precipitated in the wake of Ørsted’s discovery of electromagnetism,1 but also in the very institutions of higher learning, where more and more universities were creating separate faculties of Naturwissenschaft, and in the popularization of natural science by enthusiastic

1

Cardwell reports that the German professor H. M. Jacobi’s 1835 paper describing a rotating electromagnet with two fixed electromagnets around it and a commutator to reverse current was translated into the major European languages and attracted widespread attention, in part because of the lack in theory of an upper limit to the speed of rotation that might be produced. He writes: “A near-infinite velocity suggested a near-infinite power.” Donald S. L. Cardwell, James Joule. A Biography (Manchester: Manchester University Press, 1989), pp. 23–24.

399 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 399– 416. © 2007 Springer.

400

F. GREGORY

scientific materialists and others.2 Ørsted lived through most all of this. His death in 1851 spared him from the bulk of the controversy over materialism in the 1850s, although we have no doubt what he would have thought of it. While there are aspects of this standard narrative that have endured, it glosses over a great deal that is important. We have learned, for example, not to dismiss Naturphilosophie as an interruption in the development of “real science.” The Naturphilosoph Lorenz Oken, after all, was the motive force behind both Isis and the Gesellschaft,3 both of which were central to the beginning stages of the coagulation of a German natural scientific community. But we cannot simply buy into the idea that there was a well-established scientific community by midcentury. Recent works such as James Secord’s Victorian Sensation and Crosbie Smith’s cultural history of The Science of Energy, have underscored for us just how contested the status of natural science remained in the 1840s and beyond.4 Although both of these studies are centered on the British context, I believe the point can be made for the continent as well. But where Germany is concerned, what we have not yet done well is to get inside the process of change from Naturphilosophie to Naturwissenschaft as it happened. Ørsted may afford us insight here, because he shared sympathies with both sides in a way that few others did. John Pickstone has recently published an interesting set of reflections on the sweep of science, technology, and medicine from the Renaissance to the present. In his Ways of Knowing: A New History of Science, Technology, and Medicine,5 Pickstone depicts a few fundamental ways of reading the world that have been characteristic of natural philosophy in the past. He is, of course, careful to note that while various periods of history have featured or seemed to prefer one or another of these styles, they were all present in virtually all periods to some degree or other. Consequently, while his treatment of the 19th century bears a certain resemblance to the standard narrative referred to above, it does not at all share the positivistic outlook that drove it. The period of Ørsted’s life is a particularly interesting one since at its beginning STM, as Pickstone abbreviates science, technology, and medicine, was moving out of an allegedly dominant mode of natural history, the classifying and describing of things, into analysis, the diagnosis of the whole in terms of its parts and their relation to the whole. But by the time Ørsted died, at the middle of the 2

3

4

5

Cf. here my Scientific Materialism in Nineteenth Century Germany (Dordrecht: D. Reidel Publishing Co., 1977). Cf. also my The Mysteries and Wonders of Natural Science: Aaron Bernstein’s Naturwissenschaftliche Volksbücher and the Adolescent Einstein, pp. 23–41 in Don Howard and John Stachel, eds., Einstein: The Formative Years, 1879–1909 (Boston, MA: Birkhäuser, 2000); Alfred Kelly, The Descent of Darwin : The Popularization of Darwinism in Germany, 1860–1914 (Chapel Hill, NC: University of North Carolina Press, 1981); David Knight, “Getting Science Across,” British Journal for the History of Science, 29(196), pp. 129–138. For a recent treatment of Oken’s role as editor of Isis and of his role in founding the Gesellschaft, see chs. 4 and 5 of the Ph.D. dissertation by Heiderose Brandt Butscher, Lorenz Oken and NineteenthCentury German Romantic Science, York University, 2001. James A. Secord, Victorian Sensation: The Extraordinary Publication, Reception, and Secret Authorship of Vestiges of the Natural History of Creation (Chicago, IL: University of Chicago press, 2000); Crosbie Smith, The Science of Energy: A Cultural History of Energy Physics in Victorian Britain (Chicago, IL: University of Chicago Press, 1998). John Pickstone, Ways of Knowing: A New History of Science, Technology and Medicine (Chicago, IL: University of Chicago Press, 2001).

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 401 19th century, STM was gravitating, according to Pickstone, more toward experimenting, the control and systematic creation of new phenomena. Ørsted’s thought clearly confirms what Pickstone has to say about the move beyond natural history. As I hope to exhibit from his public lectures on the spirit in nature, Ørsted wished to distinguish natural science from natural history. Beyond that, at least where Ørsted is concerned, Pickstone’s chronology is too rigid. One reason why this is so, I believe, is that the years of the first half of the century are remarkably more fluid and unsettled than we have often acknowledged, far more unsettled than the standard narrative mentioned above allows. Among the indicators of this flux is an emerging concern with what might loosely be called scientific method. Likely the first name that comes to mind in this respect is John Herschel, whose Preliminary Discourse on the Study of Natural Philosophy appeared in 1830, and after him his countryman William Whewell. In Germany in the 1840s one began to see works in which the concern was not only to denounce Schelling’s Naturphilosophie (that had been underway in various forms for some time), but to begin to lay out how Naturwissenschaft was different in method. Matthias Schleiden wrote his methodologically oriented Grundzüge der wisseschaftlichen Botanik in 1842 and was joined later in his concern for method, of course, by the celebrated students of Johannes Müller’s Berlin School. You may wish to add several names of your own choosing to the list. From our vantage point these works had opened a can of worms. The problems these writers were tackling are enormously difficult—involving claims of induction and deduction, conjecture and refutation, probability and certainty. What I find interesting, although not surprising, is that they showed little humility in the face of these challenges. There was something new emerging just prior to and after mid-century, a conviction that experimental natural science had to be taken a whole lot more seriously than it ever had been before because it involved a method that was certain, not open to the kind of debate that characterized philosophy, including Naturphilosophie, as it was traditionally pursued. But because the confidence of these writers was premature, that is, because the issues that formed the contents of the articulations of their claims for natural science were in fact not so simple, consensus would take time to build. It is not surprising that they were confident, nor is it surprising that the formation of a “scientific community” was still in flux. Now what has Ørsted to do with all this, and even more, how are his religious views relevant? First of all, other than Jakob Fries in Germany, to whom I will refer now and then in this essay, Ørsted is among the earliest thinkers of the century to write about the issues of method I have identified as being more characteristic of the 1830s and 1840s. And his writings on the spiritualistic understanding of nature are among those where such matters are addressed. Secondly, as we all know, Ørsted just does not fit ready-made categories too well—he cuts across boundaries we have erected for others. There are certainly recognizable aspects of both the analytical and experimental modes Pickstone describes, but I do not sense that either is dominant. Furthermore, while Ørsted’s thought certainly illustrates different ways of knowing, his agenda is of course not knowing for knowing’s sake alone. Ørsted wishes to enlist natural science in a higher enterprise, one that is best

402

F. GREGORY

described as spiritual, not just in the German intellectual sense of geistig, but in a deeply personal religious sense. So, again, Ørsted may provide us with some early insights into the development of thinking about natural science that we cannot get from others. Now although I do not think Pickstone’s categories fit Ørsted in the manner Pickstone applied them to STM between 1800 and 1850, I do not think we need to treat those categories as an all-or-nothing-matter. I suggest that Pickstone’s ruminations can in fact help us understand Ørsted by giving us a way to think about his uniqueness. In fact, one way of depicting Ørsted’s singular stance in the early 19th century is to portray him as a Naturforscher who, in leaving natural history behind on his way toward becoming a Naturwissenschaftler, gives voice to both the analytical and experimental tendencies in equal measures. Unlike most figures, who gravitate more to one side than another, Ørsted is unusual because he insists on both at once.

2. Let me begin this session on Ørsted, science, and religion with what by now has become a familiar caveat. What do we mean by asking about “science” and “religion”? For some time now historians have been objecting to the almost irresistible hypostatization of “science” and “religion” into fixed and permanent entities in historical studies when we know that both words refer to changing activities that comprise whatever their practitioners at a given moment in history say they do. Surely, for example, we cannot simply take impressions about science and religion, or even about natural scientists and people of faith, that have been formulated in the 21st century and assume they accurately reflect how people in the early 19th century approached and evaluated knowledge about nature. It is in light of such warnings that Pickstone prefers to characterize interacting ways of knowing rather than freeze STM into fixed categories. His observation that natural history was waning around 1800 certainly appears to find confirmation in the Germany Ørsted visited as a young doctor. In the German states of Ørsted’s burgeoning career we do not yet have Naturwissenschaftler, only Naturforscher and Naturkenner. (I do not know what the comparable pairing in Danish would be.) Now I am as guilty as many others in having translated Naturforscher as “natural scientist” in things I have written, but I am now convinced that by doing so I not only supplied assumptions I had no right to, but that I misrepresented the self-perception of Naturforscher. In his introductory material to the wonderful recent English edition of Cuvier’s essays on fossil bones, Martin Rudwick recognizes this problem when seeking an English word that would embrace figures as different as Cuvier and Lamarck. He solves the difficulty by simply sticking to the untranslated word savant, adding a warning that it does not mean the same thing as our modern word scientist.6 6

Martin J. S. Rudwick, ed. and trans., Georges Cuvier, Fossil Bones, and Geological Catastrophes (Chicago, IL: University of Chicago Press, 1997), p. 13, n. 1.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 403 This is the kind of question that helps us get at things from the inside. Take a look, for example, at Johann Christoph Adelung’s famous Wörterbuch der Hochdeutschen Mundart from 1808 and you will not find an entry for Wissenschaftler, let alone Naturwissenschaftler. And there is, I am sure, a world of difference between Naturwissenschaftler and Naturforscher. We have Naturforscher in 1808 but only much later do we have Naturwissenschaftler. I think that the particular line Ørsted walks between the meaning of these two terms in his writings on the relationship between natural science and religion is informed by this difference. The first temptation, of course, is to assume that Naturwissenschaftler appears around the same time Whewell’s neologism “scientist” does in England and that in both cases what is signaled is the emergence of a certain kind of practitioner. I grant that this is likely. Although I have not yet been able to determine when the word Naturwissenschaftler emerges into common currency, I would not be surprised if it is not until mid-century. Where Ørsted is concerned I sense that he never adopts the term Naturwissenschaftler; for example, his term of preference in a writing of 1844 is still Naturforscher.7 But I should like to suggest that more is involved. Might not the presence of these two different terms, Naturforscher earlier in the century and Naturwissenschaftler later, tell us something about changing ways of knowing? Let us look at this a bit closer. Adelung reports three meanings of the word Wissenschaft, two of which, he notes, were going out of use in 1808 in favor of the third. First of the two older meanings was “the circ*mstance in which one knows something or has knowledge or a report of it.” (Der Zustand, da man etwas weiß, Kenntniß, Nachricht davon hat.) He gives an example: “I have no Wissenschaft of the matter.” Because Wissenschaft referred here to a state of awareness, this use had no plural form. A second meaning, also only in the singular and referring to a state or condition, and also becoming rarer according to Adelung, was “the essence of the clear and distinct concepts that one has, especially the insight into the connection of general concepts,” (der Inbegriff der klaren und deutlichen Begriffe, welche man hat, besonders die Einsicht in den Zusammenhang allgemeiner Begriffe) as in “a man of great Wissenschaft.” Here the word seems to refer to insight, an intuitive grasp not only of a specific piece of knowledge but also of its more general importance and significance. A man of great Wissenschaft here almost means a man of great wisdom, though in the special way indicated. It was the third meaning of Wissenschaft that Adelung reports was most commonly used in 1808 and was apparently displacing the other two meanings. 7

For example, the challenge for the Naturforscher is to show the truths that reflection and observation teach contain rich material for the imagination. In so doing the Naturforscher assumes the responsibility for making clear the truths gained from wissenschaftliche Naturforschung and representing the Naturleben they contain to others. Cf. the essay identified in its subtitle as submitted to the meeting of Scandinavian philosophers in Christiana in 1844, Die Naturauffassung des Denkens und der Einbildigkeit, in Der Geist in der Natur. 2 vols. (Munich: Cottaschen Buchhandlung, 1850), I, pp. 99–100. A one-volume English translation of this work, which contains several additional essays not in the German edition, appeared in 1852 and was later reprinted. The Soul in Nature, trans L. and J. B. Horner (London: Dawsons of Pall Mall, 1966). Although all translations from Ørsted’s original German edition that are given in this paper are my own, locations in the 1852 English translation are also provided as ET, here p. 43.

404

F. GREGORY

“Objective truths, [derived] from the essence of general truths that are grounded in one another; [through this] Wissenschaft distinguishes itself from art in that art contains mere practical expressions while Wissenschaft has general truths grounded in one another.” Since here the word refers to truths, it can exist in the plural. Adelung adds that “There are accordingly as many Wissenschaften as there are [ways in which] general truths, or truths of one kind, are examined as grounded in one another.” This third meaning of Wissenschaft is the one that anyone working on Kant and Schelling recognizes immediately and involves a lot of what Pickstone means by analysis, where the relation between parts and whole is paramount.the entry for wissenschaftlich, which Adelung notes comes from the third meaning of Wissenschaft, also corroborates what Pickstone suggests about the period. “Wissenschaftliche knowledge traces individual things to general concepts and understands their grounds and connections; in contrast to merely historical [knowledge], which knows only that individual things are there and at best how they are there.” As an example Adelung gives “to treat a matter wissenschaftlich, according to general concepts and propositions.”8 Clearly from the above a science of nature, a Naturwissenschaft, is implied. Indeed, there is an entry for Naturwissenschaft: It is “the wissenschaftliche knowledge of nature, i.e., of the forces that cause change in all corporeal things. Knowledge of nature, objective as well as subjective, examined as a Wissenschaft. Physics, which in other respects is also called Naturlehre, Naturkunde or Naturkenntniß.” Is there any reference to a practitioner of any sort with respect to nature? Yes, of course, but here there is no mention of Wissenschaft—of relating laws of nature to the general context in which they are embedded. This practitioner is still, in Pickstone’s categories, a natural historian. A Naturforscher (the feminine form Naturforscherin is also specifically listed) is “a person who seeks to investigate the changes in nature, according to its laws and manner of origination. Hence Naturforschung, the effort to investigate the changes bodies undergo and the laws according to which [changes] take place.” Naturkenner and Naturkennerinnen are persons “who know the changes in nature, i.e. in the corporeal world, according to their existence, their manner of origin, and the laws by which they occur.” What comes next, in its appeal to the phrase “der Inbegriff der klaren und deutlichen Begriffe” links the meaning of Naturkenner to the second, older understanding of Wissenschaft as wisdom, where almost the exact reference to clear and distinct ideas was used. Naturkenner have “knowledge of nature, i.e. the essence of the clear and distinct ideas of the general forces that cause change in bodies—also called Naturkunde.”9 These practitioners, Naturforscher and Naturkenner, correspond here much more to the merely historical knowledge—knowledge that something is or at best how it is—that Adelung contrasted to wissenschaftliche knowledge. (See the reproduction of Adelung’s definitions on the next page). Naturforscher refers mainly to practitioners as information seekers, while Naturwissenschaftler, when

8

9

Johann Christoph Adelung, Grammatische-kritisches Wörterbuch der Hochdeutschen Mundart, 4 vols. (Vienna: Pichler, 1808), IV, pp. 1582–1583. Ibid. III, p. 445.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 405 it later comes into usage, carries with it, implicitly to be sure, a much broader claim. The implication is that Naturwissenschaftler have realized that the knowledge of nature constitutes a much more fundamental place among the so-called general truths of knowledge than had ever been acknowledged before. They are Wissenschaftler, not Forscher. That, however, will not be for some time. In the

406

F. GREGORY

intervening years the term Naturforscher carries with it a certain ambiguity that allows it sometimes to refer to the natural historian, while at others to incorporate aspects of what Pickstone means by analysis and experimentation. 3. There are several places where Ørsted makes clear that his contemporaries sometimes misunderstood natural science to be natural history when he understood it to be more than that. His concern with religion provides one such context, specifically his clash with the Danish religious cleric and theologian Nikolaj Frederik Severin Gruntvig,. This exchange occurred at a time when, by many accounts, Ørsted was beginning to distance himself somewhat from Schelling. I have only been able to locate discussions of others about the exchange, the texts of which, I suspect, are only in Danish.10 Frederick Grundtvig was born into a family of old-time Lutheran Christianity. His father was a minister who showed no sign of having been influenced by Enlightenment rationalism and who wanted nothing more for his daughter and four sons than that they continue to protect the conservative heritage he himself clung to. Frederick eventually did embrace that vision, but not before being swept up as a youth in his twenties in the general enthusiasm for Schelling. As a student of theology with little inclination to natural science, Grundtvig briefly succumbed to the romantic insistence that only through a cognition informed by aesthetics could the depths of reality be portrayed. But Grundtvig soon came to the conclusion that Schelling had accomplished too much. If among the opposites reconciled in Schelling’s philosophy not only nature and spirit, but also good and evil were harmonized, then it was a castle built on air. Freed from his captivation by Schelling’s thought, Grundtvig found himself without any foundation and plunged into a deep personal intellectual and emotional crisis. So profoundly did he come face to face with a sense of his own evil that it was unclear to those around him that he would ever emerge again. When he did it was through a conversion experience that left him wary of all intellectual attempts to go beyond the biblical articulation of the historic Christian faith. The mistake of the modern age, he came to believe, was to presume that the literal word required improvement.11 Grundtvig recovered to become a noted pastor, one of Denmark’s famous theologians, perhaps its most famous writer of hymns, and the author of a vast amount of religious poetry and prose. He became known as an uncompromising defender of the faith, especially as articulated in the Bible, whose literal words he took with the utmost seriousness. 10

11

A brief summary of Grundtvig’s meaning for Denmark may be found in F. J. Billeskov Jansen, “Copenhagen—City of the Muses,” pp. 14–17 in the Ørsted commemorative special issue of Danish Journal (Copenhagen: Danish Ministry of Foreign Affairs, 1997). For an account of his flirtation with Schelling, his personal crisis, and his recovery, see Hal Koch, Grundtvig, trans. L. Jones (Yellow Springs, OH: Antioch Press, 1952), chs. 2 and 3. For an example of his defense of a literal reading of the Bible, see his What Constitutes Authentic Christianity? trans. Ernest D. Nielsen (Philadelphia, PA: Fortress Press, 1985), pp. 17 ff. This work, which originally appeared in 1826, responds to the rationalistic interpretation of scripture that predated the emergence of what is known as Higher Criticism. Cf. My Nature Lost? Natural Science and the German Theological Traditions of the Nineteenth century (Cambridge: Harvard University Press, 1992), 28 ff.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 407 But, as Llewellyn Jones has noted, Grundtvig was “a champion of old-time Lutheran orthodox piety , but with a new slant.”12 His opposition to the tendency of the educated classes to look down on their own language, to leave the use of Danish to lower classes, was consistent with his central role in the creation of a new folk school educational movement for which he is well remembered. Indeed, his role in the spiritual and national awakening of 19th-century Denmark, an awakening that showed itself in part in the establishment of industrial and agricultural cooperatives, has sometimes caused Danes to remember him as more important in the creation of modern Denmark than other well known cultural icons like Andersen or Kierkegaard.13 The immediate occasion for the dispute with Ørsted, which took place around 1814, was due to something Grundtvig had published in which he found evidence for the lack of spiritual life in the thriving of certain sciences. Earlier in his life Grundtvig had, like Ørsted, been captivated by the new German philosophy coming out of Jena. But following a religious crisis in 1810 he had become convinced that Schelling’s nature philosophy in fact made the notion of a personal God impossible and thereafter associated the flourishing of Naturphilosophie with a decline in personal spirituality. Of course the charge that Schelling’s philosophy was dangerous, that it was atheism and must be avoided, was not new. That exact warning had been given to the young Lorenz Oken at the beginning of the century by one of his medical teachers.14 We know that Ørsted certainly did not believe that the study of nature undermined spirituality; on the contrary, he was firmly convinced of the presence of spirit in nature. Further, he resented a great deal the charge that the flourishing of the sciences was a sign of a decline in spirituality. An occasion to reply presented itself to Ørsted when Grundtvig published some materials, not his own work but that of his father and a predecessor of his father. These materials had to do with biblical references to the anti-Christ, which were interpreted to point to Frederick II of Prussia by the predecessor, corrected to Napoleon by Grundtvig’s father. In the course of criticizing such interpretations Ørsted referred to Grundtvig’s “thoughtless declamations against the sciences, of which he has no idea,” adding that they had “emptied his works of all love and humanity.” In reply Grundtvig remained unconvinced that the mathematics Ørsted seemed to revere could serve at all to provide contents to an idea.15 Years later Ørsted seemed to have Grundtvig and those like him in mind when he contributed an essay to the Danish Popular Journal in which he set himself the task of defending astronomy from the accusation, which he had personally encountered, that its defense of the Copernican system was opposed to Christianity. Cast in the form of a dialogue, Ørsted’s Alfred is trying to persuade Northlight 12 13

14

15

Lewellyn Jones, Introduction, in Koch, Grundtvig, p. ix. A. M. Allchin, N. F. S. Grundtvig: An Introduction to his Life and Work (London: Darton, Longman & Todd, 1997), pp. 12–13. Oken had sketched out a system of nature philosophy in 1802 and made the mistake of sending it to one of his empirically oriented medical teachers, J. M. A. Ecker. Ecker’s response is cited in a letter of Oken to a friend: “What do you want with this mysticism? No one understands it but a few of the new nature philosophers who are despised everywhere! I can tell you, my friend, that this piece must not be printed here because everything Schellingian leads to atheism!” Letter cited in Emil Kuhn-Schnyder, Lorenz Oken: Erster Rektor der Universität Zürich (Zürich: Verlag Hans Rohr, 1980), p. 13. As reported by Jones in Koch, Grundtvig, p. xii.

408

F. GREGORY

to reexamine his dogmatic commitment to Tycho Brahe’s geocentric view of the cosmos over against the Copernican view. Northlight refuses to be impressed with Alfred’s defense of Copernicus through allusion to predictions based on the careful and patient accumulation of empirical information about the heavens. After all, that was exactly what Tycho had done. That Copernicus also could do it meant that such predictions were coincidence. Alfred distinguishes Tycho’s achievement from Copernicus’s by suggesting that Tycho’s success was founded on the collection of “some tolerably complete observations on a whole series of changes [that have been seen to] recur constantly in the same order.” Tycho then merely predicted that such observed regularity would continue. “Prophecies in this sense do not belong to science,” observed Alfred; “This kind of prediction clearly requires no insight into the laws of nature.”16 Tycho, in other words, was a natural historian.

4. Ørsted could not articulate his view of the relationship between the natural and the spiritual if scientific investigations of nature were confined to natural history. To go beyond natural history Ørsted introduces the notion of natural science as systematic knowledge. The Copernican system, he explains, is of one piece with a Newtonian world order that expresses universal and necessary laws of nature. With these laws many phenomena, including the paths of new planets, the shape of the earth, and the return of Halley’s comet, all find a common explanation. Northlight begins to waver, conceding “that there is a great unity in this doctrine.” But then he regains his composure and asserts that he cannot permit the unity to mask the fact that it is opposed to the Bible, that the Bible must be understood literally. In this piece Ørsted’s concern is to oppose the literalist mentality that Grundtvig insisted on in his attempt to protect the historic truth of Christianity. “The letter killeth, but the spirit maketh alive,” Alfred declares, quoting an unidentified biblical text.17 Yet Ørsted is sympathetic to what he is asking of Grundtvig and his followers. He knows that what he demands is in effect that believers cede to natural scientists the authority to say what is true about the realm of creation and that to do so is not easy for them. Having painted Northlight into a corner on the particulars, Ørsted has him confess that while it may be impossible for him to overcome Alfred’s arguments, still, “all my Christian feelings are opposed to the doctrine of astronomers.” All your natural science is adverse to the disposition of my mind; it transforms the whole mode of thought, and turns it away from God. In your science, it is not He who permits the sun to rise and set, or who holds the earth in His hand, or who gives it summer and winter. No; with you it is the blind laws of nature which accomplish this. It is not His anger which emits the lightning. No! With you it is only an electric spark, driven by blind necessity. It is not His power which permits the storm to sweep

16

17

“Christianity and Astronomy,” in The Soul in Nature, p. 436. This essay, indicated as taken from the Danish Popular Journal of 1837, is not included in the original German edition. Ibid. p. 443. The verse is II Corinthians 3:6.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 409 over the earth. No! It is disturbed equilibrium. It is not His goodness which sprinkles the earth with the waters of the heavens. No! It is only, as I have been told, a sport between warm and cold currents of air.18

Clearly Ørsted believes that a classic deistic argument can get him out of the hole he has dug for himself with Northlight’s eloquent objection. What he apparently never realized was that appealing to a remotely acting God would not work for most people. Ørsted has Alfred argue that if God is the sole author of the laws that govern nature, then it is literally true that “God permits the sun to rise, that he orders the change of seasons, and the course of lightning.”19 But Ørsted betrays that he suspects this argument works no better for him than it had for Lamarck (or, incidently, than it would for Robert Chambers seven years after Ørsted wrote this apology). Alfred obtains a grudging concession from Northlight that people could be good Christians whether or not they believe in the Copernican system, but Northlight then asserts that the issue may therefore be viewed with indifference. One knows that Northlight does not accept Alfred’s insistence that natural science and religion should unite in endeavoring to raise humankind above the senses, that issues such as the truth of the Copernican system are not matters of indifference because we are striving for a higher degree of spiritual enlightenment.20 Northlight has not been persuaded. Ørsted also attempted in other essays to express the relationship between his understanding of nature and its relevance to his own religious faith through a positive exposition of his philosophy of nature and science. The question we must decide is how comfortable Ørsted was with the abstract and intellectual religious sense of the German philosophy he had imbibed. Clearly he did not wish to divide things as Kant had, restricting religion to the moral realm and reserving all cognitive rights solely to Wissenschaft. Ørsted wanted to be able to unite the natural and the spiritual worlds in a manner Kant would not condone. His position came closer to that of Schelling, but it was not identical to Schelling’s, in the main because Ørsted’s way of knowing was a blend of Pickstone’s analysis and experimentation. Right around the time of the dispute with Grundtvig Ørsted delivered an address to a university gathering in commemoration of the Lutheran Reformation. It was a classic sermon based on a romantic philosophy of nature and it represents Ørsted’s sympathy for wissenschaftliche analysis, for unpacking what first appears to the casual thinker into more basic parts of a fundamental whole. When one is able to do this, when one has grounded the parts in the whole, one has confirmed their legitimacy and for Ørsted their reality. Once again, and this I believe is an important point that distinguishes Ørsted from Schelling, he carries out his exposition not merely as an exercise in philosophical analysis, but also in service of greater spiritual meaning. He ridiculed the foolish idea that “the development of the sciences (Wissenschaften) has ever endangered faith and piety,” countering it with the notion that science itself was a religious duty because research has as 18 19 20

Ibid. p. 443. Ibid. p. 444. Ibid. p. 445.

410

F. GREGORY

its ultimate goal “to find what really exists and to view it in its pure wholeness, separated from everything which deceives the careless observer by only an apparent existence.”21 The material realm, he observed, obviously consists only of that which is changing, that which is “always on the road between birth and death.”22 Elsewhere also Ørsted had labored to demonstrate how clearly chemistry through countless scientific (naturwissenschaftliche) investigations had made clear that material bodies are dissolved into misty images of air and vapor.23 Only the constant (das Beständige) was worthy of devotion. That constant, for Ørsted, resided, as he said, in “the forces that produce things and the laws by which they act.” Had he stopped here one might have assumed that he was speaking of a mechanistic conception, for the identification of the forces and their laws lay at the heart of the Newtonian dream. But how could the touch of cold philosophy evoke devotion? Ørsted followed Schelling beyond mechanism, but he then pushed on beyond Schelling to a position which for Ørsted possessed the capacity to evoke religious devotion. The powers resolve themselves into one fundamental power and the individual laws into one law of reason pervading all of nature. When one realizes and embraces this truth, Ørsted told his listeners, then one realizes “that this is not merely a concept, an abstract idea, as it is called, but that reason and the force to which everything owes its essential nature is only the revelation of an independent, living Omniscience.”24 The constant is alive, a giver of life; the constant is perfect wisdom, the generator of the harmony of the whole. All this meant something very obvious to Ørsted that was not obvious at all to either Schelling or Grundtvig. “How can he [ who seeks the constant],” declared Ørsted, “be otherwise animated than by the deepest feelings of humility, of devotion, and of love?”25 Schelling, whose constant was usually articulated in terms of the neutral Absolute, would likely have been uncomfortable with the degree of the personal Ørsted had found, while Grundtvig would hardly have been satisfied that it, not being expressed in specifically Christian terms, was personal enough. Ørsted was clearly content to go beyond Schelling’s more abstract language, but Grundtvig’s charge that the sciences detracted from spirituality apparently caught Ørsted’s attention, interrupting his beatific vision of nature sufficiently to spark some awareness that others may view things differently. Ørsted asks at the end of his address whether his conclusions meant “that all people should be men of science (Männer der Wissenschaft)?” His answer is an unconvincing no. Each must choose a particular sphere of action; however, there is no avoiding joining in Ørsted’s program of contributing to the perfection of the whole.26 One reason why Ørsted’s conclusions strike us as they do is because he is not arguing his position so much as he is declaring it. It is, in effect, a declaration of his faith. 21

22 23 24 25

26

“Die Kultur der Wissenschaft als Religionsübung betrachtet,” Geist der Natur, I, pp. 318, 320–321; ET, pp. 135–136. Ibid. I, p. 322; ET, p. 136. Cf. Das Geistige in der Körperlichen. Ein Gespräch, in Der Geist in der Natur, I, p. 11; ET, p. 5. Kultur der Wissenschaft, Geist in der Natur, I, p. 322; ET, p. 136. Later in the address Ørsted associates the contemplative existence of the divine harmony as reason while its active existence is love. Geist in der Natur, I, pp. 328–329; ET, p. 139. Ibid. I, p. 333; ET, p. 141.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 411 5. Elsewhere he does attempt to argue his case, and in a decidedly different tone that is more responsive to the demand of the empirical. In his piece on The Spiritual in the Material, written again as a dialogue among four interlocutors with Alfred once more speaking most directly for Ørsted, he sounds less like Schelling than before. Once again he asserts that we must seek what is constant, what does not change, and that one finds it in the laws of nature. But this time he specifically refuses simply to declare his belief that reason’s capacity to express the laws of nature stems from the divine wisdom revealed in nature. After all, he notes, I could be deceived in my belief. Nor, in what must be seen as an allusion to Schelling’s program in his transcendental philosophy, will he claim to deduce natural laws from the highest source of knowledge, although he appears to leave open the possibility that someday this might be possible. When asked how he will prove (erweisen) his proposition Alfred declares that he will do so “from a fact,” backtracking immediately to “from a sum of facts,” and then to “in the usual procedure of investigators (Naturforscher) of nature.” They direct their thinking to the objects of experience that are most completely known to us, and which at the same time form points of light in the scope of our knowledge, and they search for their laws.27 He will prove his case from a group of facts that reveal the connection that subsists between nature and our minds. Immediately the question arises about the validity of laws inferred. One recognizes in the structure and flavor of what follows the influence of Kant. For example, the interlocutor Hermann suggests that one could only be successful in such an endeavor if the results were expressible mathematically. While agreeing that mathematics must be an element of all perfect knowledge, Alfred resists so narrow an application of mathematics that laws like those discovered for the voltaic battery and in chemistry would not be included. Alfred’s declaration that in testing the laws we infer we are not testing the work of our own reason, but that we test the agreement of our reason with a work we are certain our reason did not produce, provokes a query from Hermann about whether this is certain. “Could not everything we regard as the external world perhaps be only the work of an unconscious activity of our own mind?”28 Ørsted clearly wants to squelch such “idealism,” as he calls it here, and he treats it in a mode that again reminds us much more of Kant’s concerns than Schelling’s. He begins by specifying what he regards as the proper starting point: that in our understanding the internal and external are not two separate things. Hermann then asks if that means he cannot assume that “in the collective outer world there is something that makes an impression upon us, but that is constituted completely differently than we imagine it to be,” so that what we call natural laws might in the end be nothing more than laws of our own manner of intuiting.29 There follows a quick lesson on the difference between sensibility and understanding in which Alfred explains that in observing some natural phenomenon 27 28 29

Geistige in der Körperlichen, Geist in der Natur, I, pp. 26–27; ET, p. 11. Ibid. I, p. 31; ET, p. 13. Ibid. I, p. 32; ET, p. 14.

412

F. GREGORY

different sentient beings may experience different sensations (for example of color) while the causal relation lying behind what is sensed is the same. He gives several examples, one of which is to compare our experience of a falling body with the experience a differently constituted resident of Jupiter might have. In the discussion Ørsted makes clear that time and space would have to be the same for such a being as they are for us, that space and time are necessary forms of the finite. Furthermore, the creature from Jupiter could not have laws of reason that are differently constituted since “there is but one reason.” Hence, what accounts for the similarity in the experiences of the two beings, the experience of the relations among sensations, is the corresponding similitude in external circ*mstances. Sensations may differ and reality itself may be different from the manner either senses it, but the laws of reason governing the mind match the laws that govern the external world. Alfred summarizes by observing: “If the laws of our reason did not exist in nature, we would strive in vain to force them on her; if the laws of nature did not exist in our reason, then we would not be able to grasp them.”30 Ørsted is clearly unwilling to leave such preestablished harmony result solely from what he calls “our thought experiment.” He delights in insisting that, while metaphysics is one route to follow, one must also begin with the constraints nature imposes. For example, it is natural to expect that there be preestablished harmony between the laws of reason and the laws of nature because humans are the product of nature.31 This observation provokes outcries of objection: Is man a mere production of nature? Should we not start with the idea of God, the original reasoner? Are you not putting too much stress on the world of the soulless, of death? In Alfred’s reply we recognize the part of Ørsted’s outlook that came from his work as an experimental natural philosopher. Alfred declares: “It is my real opinion that our discernment would be in a bad way if our living spirit did not learn from nature, which we call dead.”32 Reason executed in nature is without freedom; hence it is involuntary and infallible. But because we exercise reason with freedom, it is susceptible to error. And err we have. How inclined man is to regard himself as the midpoint of all existence! Around him the heavens are said to turn. The stars of heaven are supposed to announce his fate. For his sake everything is said to have been created. Do you believe that without knowledge of nature (Naturkenntniß) humans would have loosed themselves from these fancies? 33

It is the scientific study of nature, not cursory impressions of it, that is necessary. Elsewhere Ørsted noted that just as our earth has passed through developmental phases in its past, so our solar system has in the past and will continue in the future to submit 30

31

32 33

Ibid. p. 41; ET, 18. The same issue, with a similar use of the planet Jupiter, is dealt with in Das Ganze Daseyn ein Vernunftreich, Geist in der Natur, I, pp. 230 ff.; ET, 97 ff. “Geistige in der Körperlichen,” Geist in der Natur, I, p. 42; ET, p. 18. In “Die Naturauffassung des Denkens und der Einbildigkeit,” Ørsted contemplates the era of precivilized human life. He refers to the developmental history of the earth with its succession of eras, each with new and more perfect beings. The human race did not come into being prior to the last great transformation of the earth. Cf. Geist in der Natur, I, pp. 104–105, 122–123; ET, pp. 45, 53. Geistige in der Körperlichen, Geist in der Natur, I, p. 44; ET, p. 19. Ibid. pp. 44–45; ET, p. 19. Alfred goes on to denounce contemporary philosophers who pay no attention to natural science.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 413 to creative transformations. It would be strange indeed if none of the other planets, with their successions of created beings, had matched the earth’s degree of development. And our solar system is but one in a vast higher system, and that of a still higher one. It lies in the nature of things that reason should emerge into self-consciousness not only in one place, but in each of the members of the world system.34 It has been natural science that has corrected human error to the benefit of religion. Humans have a natural inclination to ascribe events incomprehensible to them to spirits that have human passions or they give human voluntary dispositions to God himself. Does not natural science (Naturwissenschaft) drive away many presumptions of God’s arbitrary actions, which only too often have marred godliness itself ? 35

In 1811 Ørsted wrote a sketch of what he called the essence of physics as a program of lectures, later published in Danish and then in 1822 translated into German for Gehlen’s Journal für Chemie und Physik. It remained basically the same when reprinted in Der Geist in der Natur under the title Über Geist und Studium der allgemeinen Naturlehre.36 In this essay the move toward the mode of experimentation that Pickstone identifies with the creation and control of phenomena becomes explicit. In the first section of this essay Ørsted rehearses once again the importance of seeing parts in relation to the whole in which they are embedded, emphasizing that natural science is good in and for itself and should not be valued with respect to some purely utilitarian function. But of course it will be useful by virtue of the harmony of reason that animates everything. Consequently the [excellence] of science “may be comprehended in the one great truth, it teaches us to govern nature.”37 But what follows in the second section is a treatise on the methods of natural philosophy, beginning with an emphasis on observation and experimentation, the compelling of nature to act under our eyes. What follows here reminds me of Jakob Fries’s response to Kant in the early years of the century. Fries felt that Kant, for all his excellence in establishing the metaphysical foundations of natural science, had never addressed the issues of scientific method. He saw Schelling’s Naturphilosophie as a step in that direction, one that shared his conviction that Kant had erred in his notion of organism and therefore of the impossibility of a science of living things.38 Of course Fries felt that, in spite of this agreement with Schelling over Kant, Schelling’s Naturphilosophie had turned out to be based on a fundamentally flawed foundation. Schelling had never grasped the basic point, one clearly emphasized in Kant, that there could be no such thing as an intellectual intuition 34

35

36 37 38

Die Naturauffassung des Denkens und der Einbildigkeit, Geist in der Natur, I, pp. 123–124; ET, p. 53. Geistige in der Körperlichen, Geist in der Natur, I, p. 45; ET, p. 19. In this connection see Ørsted’s essay, Überglaube und Anglaube in ihrem Verhältnis zur Naturwissenschaft, Geist in der Natur, I, pp. 131–214; ET, pp. 56–90. See comment at the beginning of the essay in the German text, Geist der Natur, II, p. 431. Über Geist und Studium der allgemeinen Naturlehre, Geist in der Natur, II, p. 445; ET, p. 452. Fries credited Schelling with this insight and frequently emphasized its importance by printing in large letters the words “NATURE IS AN ORGANIZED WHOLE.” Cf., for example, Fries’s Reinhold, Fichte und Schelling [1803] in Sämtliche Schriften (Aalen: Scientia Verlag, 1969ff), XXIV, pp. 179–180. Ørsted too identified Schelling’s genius and contribution with his seeing nature as a single organism. See quotation of Ørsted’s 1807 letter to Oehlenschläger by Andrew D. Jackson, “Introduction,” Selected Scientific Works of Hans Christian Ørsted, edited by Karen Jelved, Andrew D. Jackson, and Ole Knudsen (Princeton, NJ: Princeton University Press, 1998), p. xxv.

414

F. GREGORY

of nature. But for all that Fries acknowledged Schelling’s Naturphilosophie when he first encountered it as “the sole original great idea to show itself in Germany since the appearance of Kant’s major writings in the area of free speculation.”39 Fries’s correction of what he saw as Kant’s other mistake, an erroneous view of chemistry, reminds us of Ørsted’s. Kant taught that the only real natural science was mechanics since the laws of mechanics alone could be grounded on a priori first principles. If, as Adelung had given it, “wissenschaftliche knowledge traces individual things to general concepts and understands their grounds and connections,” then, Kant was convinced, the only individual knowledge of nature for which this could be done was our knowledge of motion and the forces that cause it. After all, any knowledge of nature that was to be grounded in first principles would have to relate directly to the very forms of sensibility through which we encounter nature—space and time—and that meant, in Kant’s famous assertion, that a description of nature qualified as genuine science (eigentliche Wissenschaft) only to the extent it could be expressed mathematically. Other knowledge of nature, for example our knowledge of chemistry and of living organisms, conformed in Kant’s mind to what Adelung had contrasted to wissenschaftliche knowledge: “merely historical [knowledge], which knows only what individual things are there and at best how they are there.” Fries and Jeremias Benjamin Richter opposed Kant’s elimination of chemistry as a genuine natural science because they believed it was possible to find mathematical expressions of chemical change.40 Ørsted agreed both that Kant provided a good starting point from which to explore the philosophy of natural science and that one needed to go farther than the great thinker from Königsberg had. He agreed that Kant had been wrong about chemistry being a merely empirical science and that part of Schelling’s genius was to see nature as a single organism. But Ørsted too feels that more needs to be said about the methods of experimental science. In the second section of the essay “On the Spirit and Study of the Gerneral Theroy of Nature” he proceeds to explain how mere observation, including the forced observations of experimentation, is not the same as possessing insight into nature. There must be a correct combination of observations, which results from an art of experience that only one who holds the whole in view possesses.41 In natural philosophy the experimental manner of proceeding is completely dominant, which is why it is called experimental natural philosophy.42 Once again the aim is to illustrate the spiritual meaning of natural 39

40

41 42

For a general examination of Fries in is relation to Schelling see my Die Kritik von J. F. Fries an Schellings Naturphilosophie, Sudhoffs Archiv, 67(1983), pp. 145–157, here p. 149. Cf. also Wolfgang Bonsiepen, Die Begründung einer Naturphilosophie bei Kant, Schelling, Fries und Hegel (Frankfurt: Klostermann), passim. Cf. my “Romantic Kantianism and the End of the Newtonian Dream in Chemistry,” Archives internationales d’histoire des sciences, 34(1984), pp. 108–123, as well as my essays “Nature is an Organized Whole: J. F. Fries’s Neo-Kantian Reformulation of Kant’s Philosophy of Organism”, pp. 91–102 in Maurizio Bossi and Stefano Poggi, Romanticism in Science: Science in Europe 1790–1840 (Amsterdam: Kluwer Academic Publishers, 1994), and “The Newtonian Vitalism of J. F. Fries,” pp. 143–155 in Guido Cimino and François duch*esneau, eds., Vitalisms from Haller to Cell Theory (Florence: Leo S. Olschki Editore, 1997). Schelling’s role is discussed in the context of his influence on Fries. Über Geist und Studium der allgemeinen Naturlehre, Geist in der Natur, II, p. 456; ET, p. 457. Ibid. II, p. 469; ET, p. 459.

HANS CHRISTIAN ØRSTED’S SPIRITUAL INTERPRETATION OF NATURAL SCIENCE 415 science, for we strive not only to discover the nature of the external world but to transfer our souls into creative activity that lets us in on the development of the ideas of things. This creative activity may employ conjectures—hypotheses—that are more or less probable and which can become certainty if all the consequences deduced from it coincide with experience.43 With the knowledge we learn from experimentation we learn the telos, why something is really so, not merely why we are convinced that something is the case.44 As mentioned above, Ørsted is among the earliest of many subsequent thinkers to concern himself explicitly with scientific method. If he reminds us of Herschel we should not be surprised at Herschel’s lavish praise of Ørsted and his achievements.45 Yet in spite of its emphasis on method, Ørsted’s purpose in writing a philosophy of natural science was different from Herschel’s and would certainly stand out from the other works on method than began to flow forth at mid-century. Herschel, whose sympathies with Christianity were well known, did not make his Preliminary Discourse an apology for his religious disposition. After mid-century Ørsted’s motivation to articulate the relationship between spirit and nature was seen by many in terms of his relation to the analysis of Wissenschaft as found in Kant and especially in Schelling, not for its insistence on the fundamental role of experimentation. That kind of analysis became more and more highly suspect because it was viewed as wedded to a speculative, not an experimental enterprise. The transition to the dominance of what Pickstone describes as an experimental mentality was increasingly evident, so much so that all forms of philosophy became suspect. Clues to the new importance claimed for experimental knowledge of nature after midcentury may be found both in assertions made about philosophy’s need to change and in criticisms of such claims by philosophers. A good deal of this controversy revolved around the issue of scientific materialism, but for all that philosophy was clearly on the defensive. Adolph Cornhill’s essay Die Philosophie als Naturwissenschaft, written in 1858, attempted to reply to critics who felt that bringing induction into philosophy as a necessary component of its foundation would undermine philosophy as queen of the sciences. For example, philosopher Heinrich Böhmer referred in 1872 to a movement that was afoot in which natural science was taking over philosophy altogether. He resented how physics and chemistry had raised themselves above their sister sciences. Ludwig Büchner had, in fact, asserted that philosophy had become more natural scientific (naturwissenschaftlicher), noting that a crisis in philosophy had been precipitated by the rise of experimental science.46 And Helmholtz’s celebration of experimental natural science, in which the aim was action, is well known.47 43 44 45 46 47

Ibid. II, p. 471; ET, pp. 463–464. Ibid. II, p. 463; ET, p. 460. See the Introduction to The Soul in Nature, pp. xvii–xviii. Scientific Materialism, pp. 145–146. Like Ørsted, Helmholtz opposed materialism. But he also railed against “the tyranny of spiritualistic metaphysics.” “In order to acquire the foreknowledge of what is coming … no other method is possible than that of endeavoring to arrive at the laws of facts by observations; and we can only learn them by induction, by the careful selection, collation, and observation of those cases which fall under the law.” “On Thought in Medicine,” in Hermann von Helmholtz, Science and Culture. Popular and Philosophical Essays, edited by David Cahan (Chicago, IL: University of Chicago Press, 1995), p. 322.

416

F. GREGORY

All this, then, occurred when natural science was clearly identifying itself as an experimental Wissenschaft. Now the point is that Ørsted is a transition figure, unlike virtually anyone else. The only contemporary figure I can compare to him is Jakob Friedrich Fries, whose outlook also must be viewed primarily in light of his religious disposition.48 But Fries never displayed the deep appreciation for Pickstone’s experimental outlook that Ørsted clearly incorporated into his thought. I do not mean to overestimate the contribution of Ørsted’s stance; in fact, to me Ørsted is most interesting as an example of one who more than most reflects the depths of the changes that were molding European natural science during the first half of the 19th century and who struggled mightily to come to grips with them. His merit is that he did not succumb to the temptation that marks most of his contemporaries. He did not opt for a one-sided response. University of Florida

48

See in this regard my “Extending Kant: The Origins and Nature of J. F. Fries’s Philosophy of Science,” which examines Fries’s views in light of his Moravian roots in Kant’s Scientific Legacy in the 19th Century (Cambridge, MA: MIT Press), forthcoming.

THE SPIRITUAL IN THE MATERIAL D. M. KNIGHT

In May 1852 Charles Darwin read the recently published English translation of Ørsted’s book, The Soul in Nature,1 and in his reading notebook recorded his verdict—dreadful.2 Most books got no comment, but the year before he had found Frank Newman’s Phases of Faith3 (an autobiographical novel of religious doubt) excellent; and this certainly tells us something about Darwin’s state of mind at the time. It also indicates the difficulty of fitting the eminent Ørsted into some kind of scientific mainstream, in his own day or since: in his early lifetime, J. W. Ritter was one of the very few with whom he was closely allied,4 in opposition to the Parisian establishment. It may seem curious that The Soul in Nature, published by Henry Bohn, was almost all there was of his writing available in English5: but less so when we remember on the one hand that (rather like Alessandro Volta’s, or later William Konrad Roentgen’s work) Ørsted’s famous researches in electromagnetism were rapidly and more fruitfully taken up by others; and on the other hand, the long tradition in the English-speaking world of natural theology, turning slowly into the rather newer and vaguer notion of the scientific sublime where such straightforward and respectable writings as Mary Somerville’s can be placed.6 Nevertheless, the middle years of the 19th century were a time when there was a good deal of translation going on, and Bohn was an important publisher of it (as well as of remainders, notably in natural history). He was described7 as being in 1850 “in the zenith of his fame,” buying and improving an estate at Twickenham; and was one of the most energetic figures in the London book world, making a

1

2

3

4

5

6

7

Hans Christian Ørsted, The Soul in Nature, with Supplementary Contributions, translated by Leonora and Joanna B. Horner, London: Bohn, 1852. F. Burckhardt et al. (ed.), The Correspondence of Charles Darwin, vol. 4, Cambridge: Cambridge University Press, 1988, pp. 488, 479, 475. F.W. Newman, Phases of Faith [1850], introduction by U.C. Knoepflmacher: Leicester: University Press, 1970; F. M.Turner, John Henry Newman: the Challenge to Evangelical Religion, New Haven, CT: Yale University Press, 2002, pp. 347–352, 613–615. D. C.Christensen, “The Ørsted-Ritter Partnership and the Birth of Romantic Natural Philosophy,” Annals of Science, 52 (1995), pp. 153–185. H.C. Ørsted, Selected Scientific Works, translated and edited by Karen Jelved, Andrew D.Jackson, and Ole Knudsen, introduction by Andrew D.Wilson, Princeton NJ: Princeton University Press, 1998, p.ix. K. A. Neeley, Mary Somerville: Science, Illumination and the Female Mind, Cambridge: Cambridge University Press, 2001, pp.40, 104 ff, 126. D. N. B., “Henry George Bohn.”

417 R. M. Brain, R. S. Cohen and O. Knudsen (eds.), Hans Christian Ørsted and the Romantic Legacy in Science, 417–432. © 2007 Springer.

418

D. M. KNIGHT

fortune from “the cheap issue of works of a solid and instructive kind,” many of which were translations. Bohn was in large part responsible for the fall of 48% in the average price of a book between 1828 and 1853, from 16/- to 8/4½; and at his retirement in 1864 his stocks of new and old books were worth £53,000 and his copyrights worth £20,000.8 His books usually came in embossed case bindings in a uniform small octavo format, with small type and narrow margins. The volumes generally contained advertisem*nts for his various series, an astonishing list which included William Paley’s Works, the Bridgewater Treatises, and William Lawrence’s Lectures; and this kind of publication indicates that Ørsted’s book was perceived by a shrewd and successful commercial publisher to be a good proposition. Bohn was also publishing, in five volumes between 1849 and 1858, a full translation of the Cosmos of Alexander von Humboldt, as well as German works of history and philosophy, including Leopold von Ranke’s History of the Popes, Barthold Niebuhr’s History of Rome, and Friedrich Schlegel’s Philosophy of History; and in 1852 a work by another Dane, Joachim Frederic Schouw’s Earth, Plants and Man. These were not specialist publications, like those of the Ray Society for example, which also commissioned translations but was a club in which members subscribed and received the books as issued.9 That Society had in 1847 published a handsome translation by Alfred Tulk of Lorenz Oken’s curious Elements of Physiophilosophy, against which Darwin caustically noted “nothing”; and also English versions of more-empirical works by other Germans and Scandinavians, though its main concern was British natural history. In 1840 the more eminent and established publisher John Murray had produced, again rather handsomely, the translation by Charles Eastlake (Royal Academician, Director of the National Gallery in London, and Fellow of the Royal Society) of the less-polemical parts of Goethe’s Theory of Colours10—there was thus a considerable contemporary interest in that German-speaking world in which Ørsted’s science was rooted. In a different German tradition, the Familiar Letters on Chemistry of Justus Liebig, for whom Naturphilosophie11 was the 19th century’s black death12 (and with whom the publisher William Francis studied13) were published in similar format (and in editions which steadily grew fatter and fatter) by Taylor and Walton. One of Liebig’s English translators, William Gregory, was Professor of Chemistry at Edinburgh (where he narrowly defeated the neo-alchemist Samuel Brown14); but he was also the translator of Karl von Reichenbach, an industrial chemist turned investigator of the mysterious auras by

8 9

10 11

12

13

14

Dictionary of Literary Biography, Detroit: Gale, vol. 106, 1991, pp. 59–62. J. Barwell-Carter and J. Hardy (eds.), Selections from the Correspondence of Dr George Johnston, Edinburgh: David Douglas, 1892, pp. 260 ff, 423. J.W. Goethe, Theory of Colours, translated by C. Eastlake, London: Murray, 1840. R.-P. Horstmann and M. J.Petry eds., Hegels Philosophie der Natur: Bezeihungen zwischen empirischer und spekulativer Naturerkenntis, Stuttgart: Klett-Cotta, 1986. W. H. Brock, Justus von Liebig: the Chemical Gatekeeper, Cambridge: Cambridge University Press, 1997, p. 67. W. H. Brock and A. J. Meadows, The Lamp of Learning: Taylor and Francis and the Development of Science Publishing, London: Taylor & Francis, 1984, pp. 105–106. J. A.Wilson, Memoir of George Wilson, Edinburgh: Edmonston and Douglas, 1860, pp. 310 ff.

THE SPIRITUAL IN THE MATERIAL

419

which some people were surrounded, and of the curious substance od or odyle15 underlying many of the phenomena of the world. Reichenbach hoped to unite magnetism, electricity, heat, light, crystallisation, and chemical attraction through their connections to the vital force: which makes Ørsted seem sober indeed. The mainstream of science is not after all as easy to perceive in these years as perhaps it seemed to be towards the end of the century when Ørsted’s great discovery could be presented as an aspect of conservation of energy. Furthermore, popular science was (and is) not always at all like what active professors of science would wish: ordinary readers seek the big picture, and have their agendas and interests, very different from those of insiders; and Bohn at least must have believed that Ørsted’s book would fall in with them. Some professors are and were, after all, very good popularizers, in tune with outsiders’ taste, sometimes to the envy and distaste of their peers. The “march of mind” or intellect, elementary education (on the monitorial system) associated with revolutionary cheapening of paper (based on chemically bleached wood pulp) and improvements in printing technology, opened a vast new market for English books; and the wide-ranging, populist, Vestiges of the Natural History of Creation, loudly denounced by sound men of science, had been one of the great publishing successes in a long series of editions from the 1840s16: all publicity is good publicity in publishing as in show business. The point that bad reviews may promote sales can also be made about theological works, for Vestiges was equally denounced from pulpits. Orthodox theodicies, or indeed devotional or systematic writings, are generally less than exciting; but natural theology was a sphere open not only to beneficed clergy of established churches, like Paley, but to ministers and laymen of very various persuasions.17 Thus Joseph Priestley, whose primary career was as a Unitarian minister rather than a chemist, advocated a Christian materialism that accorded with his view of the particles of matter as mere point centres of forces18: we had no immaterial or immortal soul (that idea being a Platonic corruption of Christianity), but awaited the miraculous resurrection of the body when the last trumpet, which would be a hideous horn to some, should sound and call us to judgement. Priestley had famously proclaimed that science was taking a new direction, which Ørsted and others would call dynamical19: Hitherto philosophy has been chiefly conversant about the more sensible properties of bodies; electricity, together with chymistry and the doctrine of light and colours, seems to be giving us an inlet into their internal structure, upon which all their sensible properties depend. By pursuing this new light, therefore, the bounds of natural science may be extended, beyond what we can now form an idea of. New worlds may open to our view, and the glory of the great Sir Isaac Newton himself, and all his contemporaries, be eclipsed, by a new set of philosophers, in quite a new field of speculation.

15

16

17

18 19

K. von Reichenbach, Researches on Magnetism, Electricity, Heat, Light, Crystallization, and Chemical Attraction, in their relations to the Vital Force, translated by W. Gregory, London: Taylor, Walton & Gregory, 1850. [Robert Chambers], Vestiges of the Natural History of Creation, [1844] edited by J. Secord, Chicago, IL: Chicago University Press, 1994. My Science and Spirituality: the Nineteenth Century and Beyond, is forthcoming, London: Routledge. J. Priestley, Disquisitions relating to Matter and Spirit, 2nd ed., London: J. Johnston, 1782. J. Priestley, The History and Present State of Electricity, 3rd ed., London, 1775, vol. 1, p. xv.

420

D. M. KNIGHT

Ørsted (who had received the Royal Society’s Copley Medal, in absentia, in 1820, W. H.Wollaston as acting President making the speech) would fit this picture of the Newton of the 19th century, though there were other claimants like A. M. Ampère, Humphry Davy, and Michael Faraday in this new set. Thus Ørsted’s popular writings and addresses had a good claim on the attention of the English-speaking public, especially because (appearing just after his death, and 30 years after his great discovery) they could be seen as a kind of testament. And half a century after Priestley’s death, mild unorthodoxy was no longer particularly alarming: the fervour of many men of science was deistic, theistic, or pantheistic20 rather than exactly conforming to church dogma. As the Glasgow professor of practical astronomy J. P. Nichol put it in his effusive and wonderfully illustrated Architecture of the Heavens21 (with allegorical pictures by David Scott, as well as splendid plates of nebulae): “above, below, around—there is God; there, his universal presence”. Forces, dynamism, and divinity were everywhere.22 The Soul in Nature may seem an odd title, very far from the mainstream empiricism of the English-speaking world; but the term comes in a poem of the highly respectable Sir John Herschel which has for its title the Baconian phrase (the epigraph for his famous Preliminary Discourse, 1830), “Man the Interpreter of Nature”23: Say! When the world was new and fresh from the hand of its Maker, Ere the first modelled frame thrilled with the tremors of life, Glowed not primeval suns as bright in yon canopied azure, Day succeeding to day in the same rhythmical march; Roseate morn, and the fervid noon, and the purple of evening— Night with her starry robe solemnly sweeping the sky? Heaved not ocean, as now, to the moon’s mysterious impulse? Lashed by the tempest’s scourge, rose not its billows in wrath? Sighed not the breeze through balmy groves, or o’er carpeted verdure Gorgeous with myriad flowers, lingered and paused in its flight? Yet what availed, alas! these glorious forms of Creation, Forms of transcendent might—Beauty with Majesty joined, None to behold, and none to enjoy, and none to interpret? Say! was the WORK wrought out! Say, was the GLORY complete? What could reflect, though dimly and faint, the INEFFABLE PURPOSE Which from chaotic powers, Order and Harmony drew? What but the reasoning spirit, the thought and the faith and the feeling? What, but the grateful sense, conscious of love and design? Man sprang forth at the final behest. His intelligent worship Filled up the void that was left. Nature at length had a soul.

Thus we do not have to go all the way back to neo-Platonism to find the world soul; it was a resonant idea in scientific Britain at the time of the Great Exhibition in the Crystal Palace, of Charles Dickens’ Hard Times and the first writings of Samuel Smiles. In 1846 when Ørsted visited Britain he had attended the British Association’s meeting in Southampton, where Herschel as past President welcomed him24 with the wish that:

20 21 22 23

24

D. M. Knight, “Higher Pantheism,” Zygon, 35 (2000), 603–612. J. P. Nichol, The Architecture of the Heavens, London: Parker, 1850, p. 219. S. Schama, Landscape and Memory, London: Fontana, 1996, pp. 247–248. John Herschel, Essays from the Edinburgh and Quarterly Reviews, with Addresses and other Pieces, London: Longman, 1857, p. 737. Several of his poems are translations from the German. Ørsted, Soul in Nature, pp. xvii–xviii, where the date is wrongly given as 1836.

THE SPIRITUAL IN THE MATERIAL

421

He could trample down, and strike for ever to the earth, the hasty generalisation which marked the present age, and bring up another and a more safe system of investigation, such as that which marked the inquiries of his friend. It was in the deep recesses, as it were of a cell, that in the midst of his study, a far idea struck upon the mind of Ørsted. He waited calmly and long for the dawn which at length opened upon him, altering the whole relations of science and, he might say, of life, until they knew not where he would lead them to. The electric telegraph, and other wonders of modern science, were but mere effervescences from the surface of this deep recondite discovery, which Ørsted had liberated.

Nine years earlier, Herschel, preparing to return from his time in South Africa, had hoped to avoid25 “the abominable speechifying & flummery of the September [1838] meeting of the British Association, and I think our Hanoverian trip may carry us nicely on the safe side of that treacly affair…there has crept into their meetings a style of mutual be-buttering the reverse of good taste.” We should thus remember that nobody is on oath when welcoming foreign visitors at conferences; but that Ørsted was a familiar, eminent, respected and by 1846a venerable figure whose writings should not be so dreadful. Herschel coming from a German family could read the language, but rather few Englishmen without such connections could read or speak German with any facility. Richard Taylor26 edited a journal, Scientific Memoirs, which ran eventually to seven volumes appearing between 1837 and 1853, consisting entirely of translations. In his preface,27 he quoted Emil Lenz’s remark about Georg Simon Ohm’s work, that “being only published in the German language, it is unknown in France and in England.” John Tyndall (an editor for volume 7) and Edward Frankland at about this time were studying for Ph.D.s in Germany, and this became the norm for ambitious chemists—who thus had to get on top of the language. But in the first half of the century, those who aspired to a university education followed a classical curriculum: in a famous legal judgement, Lord Eldon had ruled that endowed schools should not teach a syllabus their founders had not envisaged. Their sisters, on the other hand, might receive a more modern education, involving living rather than dead languages, at least as an accomplishment: and then translation may have been a way in which women, for whom the alternative would be to become a governess, might earn a respectable living. With proper fatherly pride, the surgeon-naturalist George Johnston (whose wife illustrated his works on natural history, but who got a clergyman to correct his Latin) reported to a friend on his daughter’s translation for the Ray Society28: “Although I say it who should not say it, the translation is excellently done; vigorous English, and of the right sort.” Humboldt’s writings were translated by Elizabeth Sabine, wife of Sir Edward Sabine29 who became President of the Royal

25

26 27 28 29

D. S. Evans, T. J. Deeming, B. H. Evans and S. Goldfarb (eds.), Herschel at the Cape: Diaries and Correspondence of Sir John Herschel, 1834–1838, Austin: University of Texas Press, 1969, p. 317; see also D. King-Hele (ed.), John Herschel, 1792–1871, London: Royal Society, 1992, pp. 51–66, 121–122. W. H. Brock, “The Lamp of Learning: Richard Taylor and the Textbook,” Paradigm, 2 (2001), 2–5. Scientific Memoirs, 1 (1837) iii; the set was reprinted in facsimile, New York: Johnson, 1966. Selections from the Correspondence of George Johnston, 1892, pp. 125, 357, 199. She appears (1807–1879) in Dictionary of National Biography under her husband’s name.

422

D. M. KNIGHT

Society and who shared Humboldt’s interests in terrestrial magnetism; and also (for Bohn) by Thomasina Ross and by Elise Otté30 who “lived wholly in the pursuit of knowledge” and whose studies included a course of physiology at Harvard as well as deep learning in Scandinavian languages and history, and who later in life wrote “largely for scientific periodicals.” We know a lot about Mary Ann Evans, translator of Strauss and Feuerbach, because as George Eliot she became a great novelist31; about Mary Somerville, eminent enough to have an Oxford college named after her; and Harriet Martineau, the formidable intellectual, translator of Comte and friend of Charles Darwin’s brother Erasmus. But of many translators, especially the women among them, we know too little; and Ørsted’s were Leonora and Joanna Horner, coming from a famous learned family32 who had lived for a time in Bonn. They produced a version from the German translation, dedicating it to their friend Ørsted’s daughter Mathilde, in a slightly Germanic English which probably catches well the tone of the original. It is odd how hard it is to turn German into curt Anglo-Saxon English. At the outset, the translators remark33 that “the following papers have been arranged without any reference to the different periods of time in which they were written, but as they might best serve to introduce, illustrate, or complete each other.” No heavy editorial hand has gone through to eliminate repetitions, achieve consistency or indicate development. Thus the book is an assemblage, quite different from Davy’s posthumously published Consolations in Travel,34 which has clear relationships with lectures and addresses given over the years but has been shaped by its dying author, given time to reflect, into a series of connected dialogues which he had dictated amid the sublimities of the Alps and the ruins of Rome to his companion and amanuensis, James Tobin.35 It is also very different from the one long argument from beginning to end in George Fownes’ Actonian Prize essay, Chemistry as Exemplifying the Wisdom and Beneficence of God36 brought out by the medical publisher John Churchill in the same year that he published Vestiges. Fownes was trying to do for chemistry what many other authors had done for astronomy and natural history, popularising it through natural theology: but there are problems with a science dedicated to macho intervention in and improvement upon nature, rather than to contemplation,

30 31 32

33 34

35

36

1818–1903: see her entry in Dictionary of National Biography, the 1901–1911 volume. G. S. Haight, George Eliot: a Biography, Oxford: Oxford University Press, 1968. K. Bourne and W. B. Taylor, The Horner Papers, Edinburgh: University Press, 1994; for Leonard Horner’s obituary, see Proceedings of the Royal Society, 14 (1865), v–x. His presentation copy of On the Origin of Species was given by Joanna to the Natural History Museum in London, and will be reprinted in facsimile in the series “The Evolution Debate” edited by D. M. Knight, London: Routledge, 2003. Ørsted, Soul in Nature, p. xxiii. [H. Davy], Consolations in Travel, or the Last Days of a Philosopher, London: Murray, 1830; and see my “From Science to Wisdom: Humphry Davy’s Life,” in M. Shortland and R. Yeo, Telling Lives in Science, Cambridge: Cambridge University Press, 1996, pp.103–114, also in D. M. Knight, Science in the Romantic Era, Aldershot: Ashgate variorum, 1998, pp. 283–294. J. J. Tobin, Journal of a Tour made in the years 1828–1829 through Styria, Carniola, and Italy, whilst accompanying the late Sir Humphry Davy, London: Orr, 1832. G. Fownes, Chemistry as Exemplifying the Wisdom and Beneficence of God, London: Churchill, 1844.

THE SPIRITUAL IN THE MATERIAL

423

wonder, and enjoyment.37 A world with metals, dyes, electric batteries, fertilizers, and explosives is better than one without; and yet to many the heavens and nature’s contrivances seemed better evidence for God’s oversight and provision than were our questing mind and hand—though not to Alfred Tennyson, in one of the stanzas of his great poem In Memoriam, published in 185038: I found Him not in world or sun, Or eagle’s wing, or insect’s eye; Nor thro’ the questions men may try, The petty cobwebs we have spun.

Finding the soul in nature was not perhaps an altogether straightforward business. Ørsted’s book contains dialogues; poems; addresses to scientific congresses and at university commencements; essays on religion taken very broadly, on beauty and its opposite, on science and on history; more formal lectures, some given for women39 and one originally delivered in Latin; responses to various critics; and metaphysical writing. Some were written or spoken for popular audiences, some for students, and some are academic: some seem pieces d’occasion delivered by Denmark’s most prominent man of science, playing the role of sage; while others are clearly thought out as more timeless contributions to scientific method and worldview. It is rather a rag-bag, and there must have been few who read it from cover to cover. Our title, The Spiritual in the Material, is taken from the piece the translators chose to put first in the English edition, because that seems (and perhaps seemed to them) the nearest thing to a summary of the book’s message. Here a woman, Sophia (wisdom), is interrogating three men, Alfred (harmonizer or peacemaker, the main speaker), Felix (happiness) and Hermann (mankind) about modern science, and hearing about active powers and dynamism. She learns that solid bodies have been resolved by chemistry into space filled by active powers; it is probable that gaseous elements compose all bodies; everything is in motion, nothing at rest, in a world where everything is in a state of development; apparent rest is mere equilibrium; and solid-seeming objects endure only like waterfalls, through the Heracl*tean flux of their material parts. All that is invariable are the laws of nature, founded upon Reason, which have sometimes been deduced a priori by men of science. Space and time are40 “necessary forms of the Finite; necessary forms of sense, categories of the Finite, if we may call them by such terms…If the laws of our reason did not exist in Nature, we should vainly attempt to force them upon her; if the laws of nature did not exist in our reason, we should not be able to comprehend them.” Chemists have shown how the very same elements 37

38

39 40

J. H. Brooke and G. Cantor, Reconstructing Nature: the Engagement of Science and Religion, Edinburgh: T. & T. Clark, 1998, pp. 314–346; and see my “Why is science so macho?,” Philosophical Writings, 14 (2000), 59–65, and N. G. Coley and S. A. H. Wilmot, “Chemical Industry and the Quality of Life,” C. A. Russell (ed.), Chemistry, Society and Environment, Cambridge: Royal Society of Chemistry, 2000, pp.318–349. A. Tennyson, In Memoriam, edited by S. Shatto and M. Shaw, Oxford: Oxford University Press, 1982, p. 137. Ørsted, Soul in Nature, p. xx. Ørsted, Soul in Nature, pp. 16, 18, 25, 26.

424

D. M. KNIGHT

exist in very different plants, “so that poisonous plants, and those which afford us nourishment, do not receive their essential qualities from the elements out of which they are composed, but from the manner in which these elements are arranged, i.e. by the natural thoughts, which in them are realized.” Similarly, animal species are partial realizations of the idea of the whole animal kingdom, which is part of a still more enlarged idea and so on upwards. Finally, “the higher species of animals, in their embryo condition, proceed from lower stages of developement, which are connected with those on which the inferior animals remain, and thence they traverse successive stages before they reach the end they were intended for.” The wisdom that Sophia would have acquired from this conversation would thus have included some important scientific ideas. Point atoms, centres of force, beginning perhaps with Roger Boscovich and supported by Priestley, fascinated Faraday41 and later William Thomson (Lord Kelvin)42; and were thus an important feature (or siren song) of physics for over a hundred years. Davy’s work on laughing gas and the other oxides of nitrogen,43 and then on chlorine and acidity, had established that elements do not bear properties, which must be the outcome of (electrical) forces, and arrangements of particles; and the studies of Friedrich Wöhler and Liebig on urea confirmed these notions,44 though neither of them had any time for the dynamism of Schelling’s Naturphilosophie45 where Davy’s experiments might have been seen to be anticipated in the pure thought of the philosopher. Louis Agassiz’s ideas on classification46 published in 1858 would contain thinking like Ørsted’s on species and types; while embryological development (following the great work of Karl Ernst von Baer47) and the question of “recapitulation” were much debated48—for example in Vestiges. Sophia’s education was clearly advanced; and the emphasis upon reason gave grounds perhaps for thinking that the spiritual, and not merely imponderable forces like electricity, underlay the material. The next dialogue is set in the Tuileries Gardens in Paris in 1846, and is called The Fountain. Here Alfred is again the Socrates, but this time he has only Frank to talk to; and here wonder, the impression nature makes upon us, is the main 41

42

43

44

45

46

47

48

M. Faraday, “A Speculation touching Electric Conduction and the Nature of Matter,” in Experimental Researches in Electricity, II, London: Richard & John Edward Taylor, 1844, pp. 284–293. Lord Kelvin [W. Thomson], Baltimore Lectures on Molecular Dynamics and the Wave Theory of Light, London: C. J. Clay, 1904, p. 123. H. Davy, Collected Works, edited by J. Davy, London: Smith Elder, 1839–1840, vol. 2 (reprint, Bristol: Thoemmes, 2001); D. M. Knight, Humphry Davy: Science and Power, 2nd ed. Cambridge: Cambridge University Press, 1998, pp. 28–35, 81–87. J. H. Brooke, “Wöhler’s Urea and its Vital Force: a Verdict from the Chemists,” in his Thinking about Matter, Aldershot: Ashgate Variorum, 1995, papers III and V. B. M. G. Reardon, Religion in the Age of Romanticism, Cambridge: Cambridge University Press, 1985, p. 96; F. W. J. Schelling, Ideas for a Philosophy of Nature, translated by E. E. Harris and P. Heath, introduction by R. Stern, Cambridge: Cambridge University Press, 1988. L. Agassiz, Essay on Classification, [1858], edited by E. Lurie, Cambridge, MA: Harvard University Press, 1962, p. 8. K. E. von Baer, “Philosophical Fragments,” Scientific Memoirs, 6 [Natural History] (1853), pp. 186–238. L. Nyhart, Biology takes Form: Animal Morphology and the German Universities, 1800–1900, Chicago, IL: University Press, 1995.

THE SPIRITUAL IN THE MATERIAL

425

theme. The order and unity detected by science that lies behind our appreciation of harmonious beauty; the contrast between the beautiful and the sublime; the lifeenhancing character of beauty, in what seems an anticipation of Walter Pater49; and once again the identity of the laws of nature and those of thought. It was a typical romantic hope that science, fine art, and religion, where imagination was brought to bear upon the infinite and the sublime, would be compounded into wisdom.50 Much later in the book, we find further dialogues, on beauty perceived by the eye and the ear, in painting and in music—“works of art, derived like the solar rays from an exhaustless source, everywhere diffuse life and joy.”51 The contrary view, that music is only a sensual enjoyment, was put by another character: “Intoxication is produced by music as well as by wine, though in a different manner, and when you spoke so enthusiastically of music, you spoke as a drunken man does of wine”; and so the discussion moves on to materialism. Alfred then changes the tone by introducing E. F. F. Chladni’s acoustic figures, the patterns produced by dust on tuned vibrating metal plates, and the mathematical basis of music. Geometrical beauty, in figures like the hexagon and realized in crystals, is found in harmony of tones: as “eternal reason which surely also includes an infinity of mathematical knowledge, reveals itself in the human form, so do I also see a revelation of it in the work of the composer.” Thus the conclusion is that the pleasures of art are not just those of imagination, but have a firm foundation in nature: “Let everyone who knows how to honour Nature and Reason, also reverence the Arts.” This leads into another dialogue (again this time including Sophia, now married to Herman[n]—wisdom and humanity conjoined) set in a summer house with a fine view reached after a long walk, 25 years on from their last meeting together. We learn again that repose is death; without action there can be no life; the whole body is in a state of oscillation and vibration; and “we are unanimous in the admission of internal vibrations in light and heat, which indeed act incessantly in all bodies, nor can we deny their existence in the method by which electricity is propagated, whence it again follows, that they cannot be absent in magnetism, not even in chemical effects.”52 The Beautiful is Reason, but it is also a bodily influence: one should be both a spiritualist and a materialist, and “form a clear conception of the spiritual in nature.” The last of the dialogues (all male, and not in the Danish edition) on Christianity and Astronomy, was written for a popular journal in 1837—the Copernican system apparently being under belated attack by “the hyper-orthodox party” of biblical literalists, whose (local) hero was Tycho Brahe53 and (who like Goethe in optics) saw conspiracy and dogmatism rife in modern science.54 “All existence” for Alfred “is the unceasing work of God, in which there is an impression everywhere 49

50 51 52 53

54

W. Pater, The Renaissance, edited by D. L. Hill, Berkeley, CA: California University Press, 1980, p. xxi where the critic is compared to the analytical chemist. Reardon, Religion in the Age of Romanticism, pp. 16, 102. Ørsted, Soul in Nature, pp. 326, 328, 347, 351. Ørsted, Soul in Nature, pp. 361, 370. V. E. Thoren, The Lord of Uraniborg: a Biography of Tycho Brahe, Cambridge: Cambridge University Press, 1990. Ørsted, Soul in Nature, pp. 425, 444, 446; Karl J. Fink, Goethe’s History of Science, Cambridge: Cambridge University Press, 1991, pp. 85–90.

426

D. M. KNIGHT

of his infinitely perfect reason, which never alters… but many men are desirous that God should arrange these things according to their own notions, in opposition to nature.” Modern astronomy, the realm of law, indicates that the other planets “are just as much inhabited as our own earth, although by beings who must be in some degree different from ourselves. What an incalculable variety here springs from one great fundamental thought.” Here then we are introduced to another strand in Danish religion in the time of Soren Kierkegaard and his Hegelian bishop; and to the “plurality of worlds” controversy, which was to bring William Whewell and David Brewster into further conflict in the 1850s,55 in the wake of Vestiges. Published dialogues go back to Plato as a preferred way of conveying philosophy in nuanced form, and as an activity rather than an erudition, though second-best to face-to-face discussion. Ørsted’s dialogues, which are thus in a long tradition in natural philosophy including Galileo, Robert Boyle, George Berkeley, David Hume, Davy, and Jane Marcet, do not give us a strong feeling of real characters in genuine conversation; but as a didactic mode, they work well enough to raise questions in a more palatable form than great blocks of prose. Elsewhere (in what was the second volume of the original publication) in an essay also concerned with biblical literalism and written in response to criticisms from Bishop Mynster, we find Ørsted’s poem, The Balloon: with three characters, Ernest a German antiquarian, Frankman a German naturalist, and Calchas (the Greek seer at Troy, or a searcher) an Athenian who has been educated in Germany.56 Ernest asserts that the glory of a nation and a time depends on ripeness alone, and energy of life: For th’ impoverished present Naught then remains, but strife and learned lore And the sad story of State Policies. Eden has passed from earth and left us here A weary hermitage of misery.

To which Frankman replies that “The Spirit of Invention lives and moves,” and that brotherly Love, abolishing slavery, exalts and enobles the present above the past. Calchas will not allow preference for the past or present to be a matter of taste, but runs through the intellectual achievements of ancient Greece—ending with Aristotle: Great Spirit! thou with fearless eye surveyed All nature, still by thy light we read Wisdom in ev’ry living creature’s form. Consider us aright, and thou wilt find No pause in our deep earnest search for Truth.

Great and glorious progress, social and scientific, was the great feature of the present, in the best tradition of the past. Frankman, “the man of wise and

55

56

[W.Whewell], Of the Plurality of Worlds: an Essay, London: Parker, 1853; [D. Brewster], “Of the Plurality of Worlds,” North British Review, 21 (1854), 1–44; D. Brewster, More Worlds than One: the Creed of the Philosopher and the Hope of the Christian, London, Murray, 1854; ninth thousand, 1862. Ørsted, Soul in Nature, pp. 143, 155–162.

THE SPIRITUAL IN THE MATERIAL

427

comprehensive soul”, weighed in to emphasize the search for Nature’s laws, for Reality and Truth. For Calchas The origin of power Connecting, ever present in us all, The pious soul points to th’ Eternal Ruler, He whose Omniscient Wisdom ordereth all. The watch—an image of the vast machine Which moves a world.

The moderns have added a thousand wonders of the world to the original seven: and the ancients would gladly renounce their dryads and nymphs for our understanding of the hidden course of nature—enabling the earthbound to soar higher than the eagle, and yet escape the fate of Icarus. We are in the world of Davy’s poem on eagles,57 whose memory left a type and a desire So should I wish towards the light to rise, Instructing younger spirits to aspire Where I could never reach amidst the skies, And joy below to see them lifted higher, Seeking the light of purest glory’s prize

and of Tennyson’s Locksley Hall Not in vain the distance beckons. Forward, forward let us range, Let the great world spin for ever down the ringing grooves of change

—though his hero was talking only to himself.58 With references to the right reading of the Bible (essentially as poetry rather than science); to Schiller, Novalis, and Goethe; and to Humboldt’s Cosmos, Ørsted’s essay and poem are intended to harmonize the scientific and the aesthetic modes of thought—the watch image, familiar to English-speakers from William Paley (and criticized for leading to religion without spirituality), shows Ørsted as the child of the Enlightenment, but running through the poem are the voices also of the Romantic movement, frank, earnest, and far-seeing. Among the essays in the book, one on science and religion again emphasizes God’s role as lawgiver: the laws of nature are invariable, and “the infinite wisdom of the eternal almighty God is able to guide everything without making casual alterations.”59 Again, “the more the historian understands his art, and exhibit[s] things connectedly, still more we learn from him to understand the laws by which the events of the human race and human society are directed”—though Ørsted was sure that this was not the atheists’ blind necessity, which he compared (extremes meeting) to superstition60: “Nature must be a contrivance of Reason.” And we also meet here, as elsewhere in the book, with the idea of development, a

57

58 59 60

J. Davy, Memoirs of the Life of Sir Humphry Davy,London: Longman, 1836, vol. 2, p. 157; D. M. Knight. Humphry Davy: Science and Power, 2nd ed., Cambridge: Cambridge University Press, 1998, p. 134. A. Tennyson, “Locksley Hall,” Poetical Works, Oxford: Oxford University Press, 1953, p. 96. Ørsted, Soul in Nature, pp. 177, 178, and cf. pp. 125–128. Ørsted, Soul in Nature, pp. 60, 86–87.

428

D. M. KNIGHT

progressive change from the lower to the higher.61 As we grow up, the rational does not emerge out of the irrational, but rather “conscious reason is developed out of something of which it is as yet unconscious … the reasoning powers of man must be developed by reciprocal intercourse with the exterior world.” Even language must have evolved. Animals too think, but our faculties are higher; and nature did not degenerate at the fall of man: It is certain that, before man was created, the laws of nature were the same, that matter had the same properties, and that the living beings were subject to suffering and death … I feel myself now called upon to say that our numerous investigations on the interior structure of the earth, and the laws of its development, have shown, that long before man came into the world, many great and destructive changes had taken place, in which whole species, indeed whole races of animals perished; that in those times also many animals swallowed one another, and indeed in the bones of the earliest creatures distinct marks of disease have been traced. Such are the clear proofs we possess that suffering, destruction, sickness, and death are older than the fall of man! If any part of the Bible appears to contradict this, it may undoubtedly be reconciled by a correct interpretation62; but should the contrary be the case, which I do not believe, we must leave such passages as unexplained mysteries, until a higher knowledge is attained. I leave it to dogmatists to consider how far their doctrine of sin may be regarded in every way as indisputably correct, or whether they would profit by a further investigation.

This almost reads as if it were written tongue in cheek; but while there is irony in the last sentence, we should not assume that Ørsted was mocking, or that he did not genuinely believe that, interpreted like poetry, the Bible was the guide to life. His religion was serious and momentous for him, if not orthodox; and his science, worship.63 The other great feature of the essays as of the dialogues is the dynamic science they advocate. This is a world of Baruch Spinoza’s natura naturans, not of inert natura naturata; where matter is an expression of activity, and creation a continuing process rather than a clock-making that happened in the remote past. Men of science, Ørsted thought, had begun to realize this64: chemistry was now seen to depend upon distribution of force, and upon polarity. The phlogistic theory had embraced a large number of objects within its compass; the anti-phlogistic did not include much more, though it did incorporate the gases as fundamental constituents. But the new dynamic theory65 “enlarges … the extent of chemistry far beyond its former limits. Electricity, magnetism, and galvanism now also belong to chemistry, as it appears that the very same fundamental forces which produce this effect, produce chemical effects in another form.” In an address Of the School in Life, Ørsted rejoiced that science was playing a greater part in education, and in his peroration entreated his young friends never to forget that it is our spiritual nature which renders man the image of God, and that it is science which constantly develops this divine spark within us, partly by showing us our own internal being as in a mirror, partly by keeping before our eyes the impression

61 62

63 64 65

Ørsted, Soul in Nature, pp. 182–183. A.Thwaite, Glimpses of the Wonderful: the Life of Philip Henry Gosse, London: Faber, 2002, pp. 204–227 discusses him doing just this is his Omphalos [1857]. Ørsted, Soul in Nature, pp. xiii–xiv, 134–142. Ørsted, Soul in Nature, pp. 312–320, 241. D. M. Knight, Humphry Davy, pp. 73–88, “Forces, powers and chemistry.”

THE SPIRITUAL IN THE MATERIAL

429

of the Divinity, which is everywhere manifested around us in nature … your conviction that Reason is everywhere manifested, in great as much as in small things, will lead you to trace out the secrets of nature and of the soul, where, without the light of the soul, you would not have expected them to exist; so that what appears to the uninitiated as dead matter, will to you be a living source of knowledge.

Science thus helps us to know ourselves, as well as enabling our inner light (that religious faculty prominent in various Christian traditions) to illuminate the natural world and reveal it as a realm of reason. But we can see why the Bishop might demur. Like Davy, Ørsted does not give the impression of firmly belonging to a church: and indeed ever since 1789 the French revolutionaries’ perception of churches as enemies of freedom has gained ground. Reverence for the spiritual, and for Jesus’ example, went well for Ørsted with a world in which nothing is eternal but Reason and its creative power.66 Davy had entertained Ørsted on a visit to London in 1823,67 and had met him again when touring Scandinavia on an electrochemical and chronometrical voyage in the summer of 1824, testing on a Royal Navy steamship (HMS Comet) the “protectors” of a more electropositive metal which would inhibit corrosion of the copper bottoms of warships.68 He set out69 “on the wings of hope, aided by the paddles of steam” but had a rough trip in which the protectors were knocked off in stormy seas. He reported that Ørsted “showed me his apparatus for increasing thermo-electro-magnetism, but I have some doubts as to the multiplication.” He met him again the next day with Prince Christian, who lived in a “villa very like an English country house of the second or third class”; at a party including courtiers and the now “quite blooming” Princess; where Davy was more interested in getting permission to shoot snipe. For Davy “Ørsted is chiefly distinguished by his discovery of electro-magnetism. He was a man of simple manners, of no pretensions, and not of extensive resources; but ingenious, and a little of a German metaphysician.” That last epithet, even from a protégé of Thomas Beddoes70 and friend of S. T. Coleridge, would not have been a compliment: and this general perception in Britain (though perhaps not going as far as Darwin’s) may account for the absence of interest in The Soul in Nature in the great reviews,71 the Edinburgh, the North British, the Quarterly or the Westminster. The rather sententious and repetitive character of the book may also have meant that it did not make a big impact in a decade when honest doubt was beginning to become respectable,72 and when a pantheistic spirituality was not uncommon among men of science, now ready to deny miracles and biblical literalism—there were more forceful works

66 67

68 69 70

71

72

Ørsted, Soul in Nature, pp. 290–298. J. Hamilton (ed.), Fields of Influence: Conjunctions of Artists and Scientists, 1815–1860, Birmingham: University Press, 2001, p. 16. J. A. Paris, The Life of Sir Humphry Davy, London: Colburn & Bentley, 1831, pp. 412–414. J. Davy, Memoirs of the Life of Sir Humphry Davy, London: Longman, 1836, vol. 2, pp. 188–216. T. Beddoes, Observations on the Nature of Demonstrative Evidence, London: Johnson, 1793; J. Z. Fullmer, Young Humphry Davy: the Making of an Experimental Chemist, Philadelphia, PA: American Philosophical Society, 2000. W. E. Houghton et al. (eds.), The Wellesley Index to Victorian Periodicals, 1824–1900, 5 vols., Toronto: University Press, 1966–1989. B. Lightman, The Origins of Agnosticism: Victorian Unbelief and the Limits of Knowledge, Baltimore, MD: Johns Hopkins University Press, 1987.

430

D. M. KNIGHT

being published. Perhaps because he came from a flat country, Ørsted did not like Davy, Leslie Stephen or Tyndall seem to find sublimity especially in mountains, and a fountain does for him as well as a waterfall—by the 1850s lifting up eyes to the hills was more usual among intellectuals. In an interesting essay73 on the “Hard Church” in the National Review74 the writer, Richard Holt Hutton, who was with Walter Bagehot one of the editors, commented: There is no end which a Christian writer can have in view that we hold to be so urgent and sacred as a crusade against the moral atheism of the day, and the resistance to the death of that self-centred philosophy and worship of blind nature which is fostered by the modern idolatries of beauty, and force, and law. But how are we to fight this battle? … The orthodox Church exterminated the heretics from hopelessness of their conversion. It gave up in despair the task of availing itself of the true doctrine in their hearts to introduce a truer. And this is exactly the intellectual policy of the Hard Church. Instead of rejoicing to indicate the good there is, and bringing out clearly its conflict with that which they regard as evil, they intellectually ignore the more hopeful elements that are bound up with scepticism, that they may indulge themselves in more unrestrained ferocity against it.

Novelists whose doubting or sceptical heroes are made thoroughly miserable, and churchmen who thunder from the pulpit, had provoked this particular essay, from one who himself clearly did not go all the way with Ørsted, or with the sceptics who were turning into agnostics. The National was a Unitarian journal, under the aegis of James Martineau: Hutton was one of his disciples, and at that time a keen Unitarian, son and grandson of ministers—when he changed allegiance, joining the Church of England, the review tottered and soon died. Unitarianism was famously in the Darwin family a feather-bed to catch a falling Christian75: we should not be surprised that Hutton should condemn those who sought to make a firm distinction between true Christians and the rest. But it is interesting that he should condemn just the kind of religion, or spirituality, that Ørsted had advocated in The Soul in Nature even though he believed that there was good in it. An American minister who found Bostonian Unitarianism oppressive in its orthodoxy was Theodore Parker, who in January 1859, stricken by consumption76 had to flee the Massachusetts winter for the warmth of the West Indies, expecting never to return. His intellectual autobiography reveals a man believing that his spirituality cannot be confined within the dogma of any particular church. Progressive development was an important part of his creed (if one can use that word), and religion was evolving into its absolute or natural form, now visible. But like Hutton he saw with horror that: Of late years a new form of Atheism—the ideal, once thought impossible—has sprung up; perhaps Germany is its birth-place, though France and England seem

73 74

75 76

“The Hard Church Novel,” The National Review, 3 (1856), 127–146, p. 140. Wellesley Index, vol. 3, 1979, has an essay on the National Review and identifies the authors: on Hutton see also D. N. B. [supplement]. A. Desmond and J. Moore, Darwin, London: Penguin, 1992, p. 5. Theodore Parker’s Experience as a Minister with Some Account of his Early Life, London: Watts, 1859, pp. 36–39, 64–65.

THE SPIRITUAL IN THE MATERIAL

431

equally its home. It has its representatives in America. Besides, the Pantheists tell us of their God, who is but the sum total of the existing universe of matter and of mind, imminent in each, but transcending neither, imprisoned in the two; blind, planless, purposeless, without consciousness, or will, or love; dependent upon the shifting phenomena of finite matter and of mind, finite itself; a continual becoming of this or that, not absolute being, self-subsistent and eternally the same perfection: their God is only law, the constant mode of operation of objective and unconscious force; yet is it better than the churchman’s God, who is caprice alone, subjective, arbitrary, inconstant, and with more hate than love. … I have simply referred them to the primal instincts of human nature, and their spontaneous intuition of the divine, the just, and the immortal; then, to what science gathered from the world of matter, and the objective history of man in his progressive development of individual and of social power.

He concludes with a side-swipe at those who “with their Bibles laid humanity flat” in commercial Cincinnati, Philadelphia, New York, and Boston by handing over runaway slaves while atheists and pantheists set them a better example. He and Hutton show that adherents to the religion of reason, nature, law and force might disagree about doctrines among themselves, even if they resolved like Parker not to use harsh words as far as possible. The term “agnostic” was just coming in77 as a respectable and undogmatic alternative to “atheism”, which still strongly conveyed overtones of immorality (Hutton’s “moral atheism”): but agnostics also differed quite widely among themselves, while like Parker, Hutton, or Ørsted seizing the moral high ground from orthodox clergy. We began with doubt and dreadfulness, and have been looking at the rather attenuated religion of The Soul in Nature and the no-longer-fashionable dynamic science popularized in it. By the 1850s also the polymath was no longer much admired; the age of expertise and specialisation was upon us. Perhaps to escape from the woolliness of romantic nature worship, harder-church theologians in Germany moved away from the God of Nature.78 Men of science and theologians, especially when writing popularly, seek to dispel doubt. We all know that Committees on the Public Understanding of Science are founded on the hypothesis that if everyone knew about science, they would love it: this is clearly an illusion, but a similar spirit is found among other kinds of evangelists. In fact, people in the voyage of their lives have different requirements, and pick up or drop all sorts of tenets and activities of science or of religion (some deemed central), whether they join a church or scientific society, remain in it, or leave it. As popular science and religion idiosyncratically compounded, Ørsted’s book could not please everyone. The mass of mankind probably demand a faith less intellectual and more practical. What then can be said about The Soul in Nature as a piece of natural theology? It is not detailed, like Paley, or the Bridgewater Treatises, or George Fownes: the argument is not carried through in the cumulative way necessary for the rhetoric of Design. It lacks the effusive fervour of Samuel Parkes in the notes to his Chemical

77 A. Pyle (ed.), Agnosticism: Contemporary Responses to Spencer and Huxley, Bristol: Thoemmes, 1995, pp. xii–xix. 78 F. Gregory, Nature Lost? Natural Science and the German Theological Traditions of the 19th century, Cambridge, MA: Harvard University Press, 1992.

432

D. M. KNIGHT

Catechism79; and it could not be mistaken for straightforward popular science, in which natural theology was often the sugar to make the medicine go down. Nor is it critical of the tradition, like George Wilson’s refreshingly honest Religio Chemici [1862] though that like Ørsted’s is a series of essays rather than en extended argument— readers will not forget Wilson’s remark80 that God has been very kind to the shark: that there is “shade” as well as “sunshine.” Optimism and benevolence reign in Ørsted’s writing: which is perhaps after all best seen as on the borderline between religion and irreligion: an interesting place to be in the 1850s provided one were not satisfied with rather bland conclusions, as sometimes he is. And that category with which we began, the scientific sublime, might be a useful key with which to unlock the kind of spirituality we find there, and also in writings of Davy, John Tyndall, and James Glaisher.81 Reverence and awe, reason, and progress, are key words: like a mighty army moves the Church Scientific, but solitariness would be the characteristic of the spiritual experiences of its soldiers. University of Durham

79

80 81

S. Parkes, The Chemical Catechism, with Notes, Illustrations, and Experiments, 3rd ed., London: Lackington Allen, 1808, p. 53 note. G. Wilson, Religio Chemici: Essays, London: Macmillan, 1862, pp. 43, 49. D. M. Knight, Zygon, 35 (2000), p. 609–611.

History of Science Department Harvard University Science Center Cambridge, MA 02138

SYMPOSIUM ON H. C. ØRSTED AND THE ROMANTIC LEGACY (FRIDAY, MAY 10, 2002) 8:30 am

Greetings

Session 1: Kant (Chair & comments: Robert Brain [Dept. of Hist. of Sci., Harvard]) 8:40–9:10 Paul Guyer [Philosophy Dept., Univ. of Pennsylvania] 9:10–9:40 Michael Friedman [Hist. & Phil. of Sci. Dept, Indiana Univ.] 9:40–10:10 Keld Nielsen [Danmarks Museum, Denmark] 10:10–10:20 Questions from audience 10:20–10:40

Coffee break

Session 2: The Danish Context (Chair & comments: Ole Knudsen [Hist. of Sci. Dept., Univ. of Aarhus]) 10:40–11:10 Karen Jelved [Copenhagen, Denmark] 11:10–11:40 Anja Skaar Jacobsen [Hist. of Sci. Dept., Univ. of Aarhus] 11:40–12:40 Roundtable: Dan Ch. Christensen [Kvanløse Havremark, Denmark] Keld Nielsen [Danmarks Museum, Denmark] Andrew Wilson [History Dept., Keene State College] 12:40–12:50 Questions from audience 12:50–2:00

Lunch

Session 3: Links to German Science/Philosophy (Chair & comments: Gerald Holton [Harvard]) 2:00–2:30 Lorraine Daston [Max-Planck-Institut für Wisenschaftsgeschichte, Berlin] 2:30–3:00 Robert Brain [History of Science Dept., Harvard] 433

434 3:30–4:00

SYMPOSIUM

4:00–4:30 4:30–5:00 5:00–5:10

Ernst Hamm [Sch. of Analytic Studies & Info. Technology, York Univ.] Frederick Beiser [Philosophy Dept., Syracuse University] Robert Richards [Div. of Social Sciences, University of Chicago] Questions from the audience

5:10–5:50

Viewing Equipment at Historical Instrument Collection

(SATURDAY, MAY 11, 2002) Session 4: Spirituality/Religion (Chair and comments: John Murdoch [Dept. of Hist. of Sci., Harvard]) 8:30–9:00 Frederick Gregory [Dept. of History, Univ. of Florida] 9:00–9:30 Dan Ch. Christensen [Kvanløse Havremark, Denmark] 9:30–10:00 Andrew D. Wilson [History Dept., Keene State College] 10:00–10:30 David Knight [Philosophy Dept., Univ. of Durham, UK] 10:30–10:40 Questions from audience 10:40–10:50

Coffee Break

Session 5: Links to France (Chair and comments: Olivier Darrigol [Center for Hist. of Sci., Paris]) 10:50–11:20 Christine Blondel [Centre de recherche en histoire des sci. et des techniques, Paris] 11:20–11:50 Michael Dettelbach [Boston University] 11:50–12:00 Questions from audience 12:00–1:00

Lunch

Session 6: Links to England (Chair and comments: T.B.A.) 1:00–1:30 Trevor Levere [Inst. for Hist. & Phil. of Sci. & Tech., Univ. of Toronto] 1:30–2:00 Gordon McOuat [Dibner Institute, MIT] 2:00–2:10 Comments by David M. Knight [Philosophy Dept., Univ. of Durham, UK] 2:10–2:20 Questions from audience Session 7: Instruments and Experiments (Chair and comments: Erwin Hiebert, Hist. of Sci. Dept., Harvard]) 2:20–2:50 Olaf Breidbach [Friedrich-Schiller-Universtät Jena] 2:50–3:20 Kenneth L. Caneva [Dept. of History, Univ. of North Carolina] 3:20–3:50 Ole Knudsen [History of Science Dept., University of Aarhus] 3:50–4:20 Roberto de Andrade Martins [Group of Hist. & Theory of Sci., Campinas]

SYMPOSIUM 4:20–4:50 4:50–5:20 5:20–5:25 Adjournment

435

Heinz-Otto Sibum [Max-Planck-Institut für Wissenschaftsgeschichte, Berlin] Maria Trumpler [Harvard University] Questions from audience

INDEX

A Aarland, L. J., 178 Abich, R. A., 388–390 Abildgaard, P. C., 104 Adelung, Johann Christoph, 403, 404, 414 Amalia, Anna, 179 Ampère, André-Marie, vii, 15, 133, 309, 310, 318, 323, 324, 346, 420 Andersen, Hanne, xi, xix, 97–112 Andersen, Hans Christian, 17 Arago, Dominique François, 15, 346 Arnim, Ludwig Achim von, 105, 341, 342, 379 August, Carl, 177, 179 B Baader, 105 Bacon, Francis, 221 Baer, Karl Ernst von, 424 Bagehot, Walter, 430 Banks, Joseph, 259, 266, 267, 356 Bastholm, Christian, x, 5–10 Batsch, August Johann Georg Carl, 177–179, 186, 187, 192, 194, 195, 197, 199, 200 Batteaux, C., 123 Baumgarten, A.G., 98, 99 Beddoes, Thomas, vi, xiv, 259–272, 429 Beiser, Frederick, xvii, 59, 142, 155 Berkeley, George, 426 Berthollet, C.-L., 249, 250, 252 Bertuch, Friedrich Justin, 179 Berzelius, Jöns Jacob, 174, 259–261, 351 Biot, Jean-Baptiste, 15, 249, 251, 252, 294, 295, 299, 304, 331, 350

Black, Joseph, 262, 263, 265 Blondel, Christine, xvii Blumenbach, J.F., 99, 177, 266, 271 Bode, Johann Joachim Christoph, 179 Böhmer, Heinrich, 415 Bonaparte, Jerôme, 254 Boscovich, Roger, 424 Boscowich, R., 120 Böttiger, Karl August, 179, 254 Boyle, Robert, 390, 392, 393, 398, 426 Brahe, Tycho, 408, 425 Brain, Robert Michael, vi, viii, xiii, xvii, xix Breidbach, Olaf, vi, xii, xix, 177–212, 434 Brentano, Clemens, 179, 341 Brentano, Sophie, 179 Brewster, David, 311, 353, 426 Brooke, John Hedley, 34, 35, 423, 424 Broussais, Joseph Victor, 62 Brown, Samuel, 418 Brugnatelli, Luigi, 296, 299, 300 Buch, Leopold von, 169, 174, 175 Buffon, G.-L. Leclerc Comte de, 99, 162 Bugge, Thomas, 103 Burnet, Thomas, 162 C Caneva, Kenneth L., xiv, xv, xix, 97, 154, 156, 160, 219, 227, 273–331, 376, 434 Canton, John, 388, 390–392 Cavendish, Georgiana, 267 Chladni, Ernst, xii, xiii, 14, 122, 124, 125, 127, 128, 133, 231, 236–238, 244, 292, 425

437

438

INDEX

Christensen, Dan Charly, xii, xix, 28, 30, 115–133, 145, 150, 156, 160, 218, 220, 305, 380 Colding, Ludvig August, 160, 330, 395 Coleridge, Samuel Taylor, xiv, 66, 138, 160, 259, 262, 263, 265–272, 429 Cornhill, Adolph, 415 Cotta, J.G., 249, 251 Cronstedt, Axel Frederik, 163, 266 Cruickshank, William, 348 Cuvier, Georges, 160, 161, 172, 175, 249, 256–258, 402 D d’Alembert, Jean, 124, 128 Darwin, Charles, xiv, 182, 235, 400, 417, 418, 422, 429, 430 Daston, Lorraine, xiii, xix, 235–246, 433 Davy, Humphry, xiv, 29, 30, 138, 156, 259–262, 268–271, 420, 422, 424, 426–430, 432 de Carbonnières, Ramond, 252 de Fontenelle, Bernard, 237 de Morveau, Louis Bernard Guyton, 276, 305 de Prony, Riche, 249 de Sacy, Silvestre, 249 Delambre, Joseph, 249, 258 Desormes, Charles Bernard, 276, 374 Dettelbach, Michaelxiv, xix, 161, 247–258, 434 Doebereiner, Johann Wolfgang, 179, 197 Doederlein, Johann Christoph, 264 Dulk, Friedrich, 349 Dundas, Henry, 267 E Eastlake, Charles, 418 Eckermann, Johann Peter, 179 Einsiedel, Friedrich Hildebrand von, 179 Elias, Norbert, 51 Eliot, George, 422 Eriksson, Gunnar, 261 Erman, Paul, 275, 345, 379, 380 Erxleben, Johann Christian Polykarp, 193, 195, 196 Eschenmayer, C.A., 110, 377 Euler, Karl-Joachim, 124, 128, 294

F Falk, Johannes Daniel, 179 Faraday, Michael, 30, 34, 38, 117, 119, 128, 129, 135, 138, 156, 310, 319, 350, 354, 420, 424 Feldbæk, Ole, 24, 26, 27 Fernow, Carl Ludwig, 179 Fichte, Johann Gottlieb, xi, 21, 28, 31, 32, 40, 41, 50, 56, 101, 128, 130, 131, 141, 178, 179, 217, 221, 223, 271, 357, 362, 413 Fischer, Johann Carl, 182, 187, 193, 194, 196 Fourcroy, A.-F., 304 Fourier, Joseph, 48, 310, 329 Fowler, Richard, 340, 355 Francis, William, 418 Frayn, Michael, 237 Frege, Gottlob, 181 Frercks, Jan, 184, 193–196, 212 Fresnel, Augustin, 346, 352 Freud, Sigmund, 235, 236 Friedrich, Carl, 179 Friedrich, Caspar David, 225 Fries, Jakob Friedrich, 181, 210, 211, 401, 413, 414, 416 Fritsch, Jakob Friedrich von, 179 Fröbel, Friedrich Wilhelm August, 179 Frommann, Carl Friedrich Ernst, 179, 198, 339 Froriep, Ludwig Friedrich von, 179, 187, 199 Fuchs, Georg Friedrich Christian, 187, 192, 194, 199 G Gall, Franz Joseph, 55–62, 66, 128 Gautherot, Nicolas, 294, 301 Gay-Lussac, 249 Giddy, Davies, 263, 267 Gilbert, Davies, 263 Gilbert, Ludwig Wilhelm, 285, 288, 296–298, 300, 305, 307, 309, 347, 379, 380 Glaisher, James, xvi, 432 Gmelin, Leopold, 352, 390 Göchhausen, Luise von, 179 Gode-von-Aesch, Alexander, 365, 366 Goethe, Johann Wolfgang von, ix, 50, 160, 161, 173, 175, 177–179, 183, 197, 198,

INDEX 200, 203, 212, 222, 233, 244, 248, 271, 363, 418, 425, 427 Golinski, Jan, 29, 43 Göttling, Johann Friedrich August, 186, 187, 192, 194 Gower, Barry, 137, 146, 150, 153, 156, 160, 219 Gräbner, Karl, 179 Granet, Marcel, 364, 365 Gregory, Frederick, xvi, xvii, xix, 247, 399–416, 431 Gregory, William, 418, 419 Gren, F.A.C., 103, 116, 195, 359 Gries, Johann Diederich, 179 Griesbach, Johann Jakob, 179 Grundtvig, N.F.S., 3, 15, 16, 23, 34–37, 44, 45, 47, 48, 247, 248, 406–410 Guyer, Paul, xi, xix, 75–96, 108, 121, 132 Gyllembourg, T.C., 65 H Hachette, Jean Nicholas Pierre, 374 Hacking, Ian, 221, 379 Hamm, Ernst, xii, xix, 159–175 Hamman, J.G., 226 Hansteen, Christopher, 311, 347–349, 353, 380 Hartley, D., 261 Hauch, A.W., 102–104, 116, 122, 125, 388 Hauch, C., 22, 25, 31–35, 37, 38, 46, 48 Hegel, Georg Wilhelm Friedrich, 2, 123, 136, 141, 177, 179, 181, 208, 210, 217, 414, 418, 426 Heiberg, Johann Ludvig, 63, 65 Helmholtz, Hermann von, xiii, 123, 136, 137, 156, 236, 242–246, 415 Herder, Johann Gottfried, 5, 178–180, 226, 271 Herholdt, Johann Daniel, 60 Hermbstädt, Sigismund Friedrich, 276, 288 Herschel, John, xvi, 11, 119, 128, 227, 401, 415, 420, 421 Herschel, William, 288–292, 355 Herz, Henriette, 30, 193, 220 Hessenbruch, Arne, x, xi, xix, 21–54 Heyne, Christian Gottlob, 253, 264 Heynig, Johann Gottlob, 179 Hildebrandt, T.M., 105 Hisinger, W., 260, 261

439

Holberg, Ludvig, 6 Hölderlin, Friedrich, 179 Holton, Gerald, xvii, 433 Homer, 245, 257 Howitz, Frantz Gotthard, 43, 62, 63, 65 Hufeland, Christoph Wilhelm, 177, 179, 198 Hume, David, 62, 75–79, 86, 96, 266, 426 Hutton, James, 162, 172 Hutton, Richard Holt, 430, 431 I Iffland, August Wilhelm, 179 Ingemann, B.S., 48 J Jackson, Andrew D., vii, x, xix, 11, 13, 21, 99, 101, 106, 109, 115, 117, 160, 218, 218, 222, 230, 231, 238, 247, 260, 262, 413, 417 Jacobsen, Anja Skaar, x, xi, xix, 55–68, 100, 260, 276, 281–285, 314, 319, 358, 359, 362, 363, 369, 376, 378, 380, 387, 397, 433 Jagemann, Christian Joseph, 179 Jelved, Karen, vii, x, xix, 11, 13–19, 21, 99, 101, 106, 109, 115, 117, 137, 145, 146, 151–154, 160, 218, 219, 222, 230, 231, 238–240, 260, 413, 417, 433 Johnston, George, 418, 419, 421 Jones, Llewellyn, 406, 407 K Kant, Immanuel, vii, x, xi, xii, 1, 21, 28, 31, 41, 43, 50, 56, 62, 75–77, 79–98, 100–112, 115–118, 120, 121, 123, 128, 131–133, 135–157, 160, 163–165, 180, 210, 217–219, 221, 223, 224, 239, 240, 245, 248, 250, 255, 256, 261, 266, 269, 271, 314, 362, 376–378, 380, 381, 399, 404, 409, 411, 413–416 Kastner, Karl Wilhelm Gottlob, 187, 192, 349, 350 Keil, Johann Georg, 179 Kepler, Johannes, 133, 246 Kierkegaard, Soren, xii, xvi, 2, 14, 46, 48, 124, 128, 407, 426 Kirmmse, Bruce, 2 Kirwan, Richard, 348 Klauer, Martin Gottlieb, 179

440

INDEX

Kleist, Heinrich von, 241, 242 Klingemann, Ernst August Friedrich, 179 Klopstock, Friedrich Gottlieb, 226 Knebel, Carl Ludwig von, 179 Knight, David M., xvi, xvii, xix, 160, 184, 197, 211, 417–432, 434 Knudsen, Ole, vii, viii, xv, xvii, xix, 11, 13, 21, 99, 101, 106, 109, 115, 117, 160, 218, 219, 222, 230, 231, 238, 260, 387–398, 413, 417 Koestler, Arthur, 246 Kornerup, Bjørn, 7 Kotzebue, August von, 179 Kraft, Jens, 6 Kraus, Georg Melchior, 179 Krause, Karl Christian Friedrich, 208, 210 Kuhn, Thomas S., 136, 160, 219, 246 L Lacoue-Labarthe, Phillipe, 225 Langlès, L.M., 249 Laplace, P.S., 35, 245, 249, 252 Latreille, P.A., 249 Lavater, Johann Caspar, 226, 229, 233 Lavoisier, A., 100–102, 120, 121, 140, 147, 148, 155, 162, 166, 184, 195, 262, 263, 277, 282 Lawrence, William, 418 Lehot, C.J., 346 Lenz, Emil, 421 Lenz, Jakob Michael Reinhold, 179, 188 Lenz, Johann Georg, 179, 187, 200 Letronne, J.-A., 250 Levere, Trevor, xiv, xix, 66, 160, 259–272, 434 Lichtenberg, C.G., xiii, 119, 124, 125, 128, 163, 180, 193, 229, 236, 238, 359, 368 Liebig, Justus von, 210, 418, 424 Locke, John, 261 Loder, Justus Christian, 177, 179, 198, 199 Lowth, Robert, 264 Luther, Martin, x, xiv, 5, 26, 33, 35, 37, 44, 257, 406, 407, 409 Lyell, Charles, 169 M Malpighi, M., 99 Manthey, Ludvig, 29, 115, 116, 275, 276, 285, 294–296, 298, 301, 305, 327 Marcet, Jane, 426

Martensen, Hans, 1–3 Martineau, Harriet, 422 Martineau, James, 430 Martins, Roberto de Andrade, xv, xix, 317, 339–381 Marum, Martinus van, 120, 296, 299 Mary Ann Evans. See George Eliot Maschmann, Hans Henrik, 347–349, 352, 353, 380 Mayer, Johann Tobias, 193–195 McOuat, Gordon, xiv, xvii, 434 Meiners, Christoph, 264 Mereau, Sophie, 179 Mesmer, Franz Anton, 340, 341 Meyer, Johann Heinrich, 179 Meyer, Kirstine, 4, 31–33, 37, 39, 47–49, 103, 126, 285, 288, 294, 319, 323, 348, 357 Michaelis, Johann D., 264, 269 Morichini, Domenico, 307 Mozart, Wolfgang Amadeus, 127, 241, 246 Muller, J., 333 Murray, John, 268, 346–349, 418, 422, 426 Musäus, Johann Carl August, 180 Mynster, Jacob Peter, 3, 62, 63 N Nancy, Jean-Luc, 225 Napoleon, xiv, 38, 175, 247, 248, 250, 251, 254–258, 304, 379, 407 Nauche, Jacques-Louis, 294, 295, 301 Neiiendam, Michael, 6, 7 Newman, Frank, 417 Newton, Isaac, 104, 110, 111, 115–118, 120, 133, 135–137, 139, 147, 155, 221–223, 317, 408, 419, 420 Nichol, J.P., 420 Nicholson, William, 145, 292, 343 Niebuhr, Barthold, 418 Nielsen, Keld, xi, xix, 97–112, 290, 310, 312, 433 Novalis, 31, 161, 169, 180, 217, 218, 220, 221, 224, 225, 227, 229, 271, 363, 427 O Oehlenschläger, Adam, 3, 8, 13, 14, 123, 220, 413 Oken, Lorenz, 180, 182, 183, 197, 199, 208–211, 217, 339, 400, 407 Oldenburg, Christian, x, 5

INDEX Ørsted, Anders Sandoe, 220 Ørsted, Hans Christian, vii, ix, xi, xii, 1, 3, 5, 6, 9–11, 13–15, 17, 21–26, 28–40, 44–53, 55, 65, 69, 98–101, 104, 106, 115, 117, 119–122, 135, 137, 145, 146, 151, 152, 154, 159, 160, 170, 218–220, 222, 230, 231, 236–240, 260, 261, 273, 339, 343, 346–348, 353, 357, 358, 387, 399–417 Otté, Elise, 422 Otto, Carl, 55, 59–62, 65, 67, 71 P Paley, William, 34, 35, 418, 419, 427, 431 Parker, Theodore, 420, 426, 430, 431 Pater, Walter, 425 Paul, Jean, 121, 132, 179, 271 Paulus, Heinrich Eberhard Gottlob, 180 Pawlowna, Maria, 179 Perkins, Jacob, 390–392 Pfaff, Christoph Heinrich, 60, 270, 312, 324 Pickstone, John, 400–402, 404, 406, 409, 413, 415, 416 Pictet, Marc-Auguste, 252, 253 Plato, 50, 123, 128, 237, 257, 261, 426 Prechtl, Johann Joseph, 321, 323 Priestley, Joseph, 29, 30, 120, 145, 359, 419, 420, 424 R Ranke, Leopold von, 418 Reichenbach, Karl von, 418, 419 Reimarus, H.S., 264, 265 Reinhold, Carl Leonhard, 180 Rendu, Louis, 347, 349, 350 Richards, Robert J., xvii, 160, 163, 182, 222, 434 Richter, Jeremias Benjamin, 275, 414 Riemer, Friedrich Wilhelm, 180 Ritter, Johann Wilhelm, ix, xii, 164, 180, 185, 198, 217, 218, 229, 261, 270, 281, 339, 356, 359, 361, 372, 374 Rochlitz, Johann Friedrich, 180 Roentgen, William Konrad, 417 Rose, Valentin, 275 Rosen, Charles, 225–228, 233 Rosenstand-Goiske, Peder, x, 6, 7 Ross, Thomasina, 422

441

Rousseau, Jean-Jacques, 124, 127, 128 Rudwick, Martin, 161, 168, 169, 402 Runge, Phillip Otto, 225 S Sabine, Edward, 421 Sabine, Elizabeth, 421 Sadler, James, 263 Schelling, Frederick, xi, xii, xvi, 1–3, 21, 28, 31, 35, 55, 57, 58, 66, 110, 123, 128, 130, 131, 135–138, 141–146, 149–156, 159–162, 164–166, 170, 171, 173, 177, 178, 180–183, 185, 187, 195, 197, 207–211, 217, 219, 221, 224, 229, 239, 240, 245, 250, 251, 260, 261, 269, 271, 307, 314, 315, 320, 327, 339, 354, 357, 358, 361–364, 366, 369, 376–379, 399, 404, 406, 409–411, 413–415, 424 Scherer, Alexander Nicolaus von, 56, 105, 192, 194, 341, 378 Schiller, Friedrich von, 130, 178, 180, 223, 248, 269, 271, 427 Schimmelmann, Heinrich Ernst, 26, 34, 39, 46, 48, 60 Schlegel - Schelling, Caroline, 180 Schlegel, Dorothea, 180 Schlegel, Friedrich, 128, 180, 217, 218, 220, 221, 223–228, 233, 286, 291, 357, 418 Schleiden, Matthias Jacob, 181, 183, 203, 210, 211, 401 Schmid, Carl Christian Earhard, 180, 187 Schopenhauer, Johanna, 180 Schouw, Joachim Frederic, 49–51, 418 Schulze, Friedrich Gottlob, 180 Schumann, Robert, 217, 225, 227, 233 Schuster, Johann Constantin, 283, 287 Schütz, Christian Gottfried, 180 Schwabe, Johann Samuel Gottlob, 180 Schweigger, Johann Salomo Christoph, 318, 329, 349, 390, 392, 393 Secord, James, 400, 419 Seebeck, Thomas Johann, 306 Shanahan, Timothy, 97, 98, 150, 151, 155, 156, 219, 239, 262 Sibbern, Frederik Christian, 60, 62–68, 73, 74 Simon, Paul Louis, 275 Smith, Crosbie, 400 Socrates, 237, 257, 424

442

INDEX

Somerville, Mary, 417, 422 Spurzheim, Johann Gaspar, 55–57, 60–62, 66 Staël, Germaine de, 247–249, 254–256 Stäpfer, Philippe Albert, 256 Stauffer, R.C., 97, 135, 136, 146, 149, 160, 219, 286 Steffens, Henrik, xii, xiii, 3, 159–175, 180, 185, 369, 375, 376 Stein, Charlotte von, 180 Strickland, Stuart, xvii, 229, 230, 233, 261, 379 Sturgeon, William, 30 Suckow, Laurenz Johann Daniel, 186–188, 190, 192–196 Sue, Pierre, 340 T Taylor, Richard, 421 Tennyson, Alfred, Lord, 423, 427 Thénard, L.-J., 304 Thomson, William, 424 Tieck, Ludwig, 180, 217 Tobin, James, 422 Tode, J.C., 99 Trumpler, Maria, xvii, 219, 435 Tulk, Alfred, 418 Tyndall, John, xvi, 421, 430, 432 V Van den Bosch, 99 Van Swinden, J.H., 296, 340 Villers, Charles de, 253–255, 257 Voigt, Christian Gottlob von, 180 Voigt, Johann Heinrich, 186, 187, 295 Volta, Alessandro, 299, 301, 356 von Haller, Albrecht, 264 von Humboldt, Alexander, ix, xiv, 169, 173, 174, 178, 179, 185, 218, 248, 250, 251, 340, 341, 418 von Humboldt, Wilhelm, 179, 249, 251

Voß d. J., Johann Heinrich, 180 Voß, Johann Heinrich, 180 Vulpius, Christian August, 180 W Wagner, Michael, 26, 38, 39, 50 Watt, James, 262 Weiss, Christian Samuel, 318, 397 Werner, Abraham Gottlob, 161, 163, 167–170, 173–175 Wetzlar, Gustav, 350 Wheatstone, Charles, 128 Whewell, William, 394, 401, 403, 426 Whiston, William, 162 Wieland, Christoph Martin, 180 Wieland, Ludwig, 180 Williams, L. Pearce, 97, 129, 135–138, 156, 219 Wilson, Andrew D., vii, x, xix, 1–10, 115, 137, 160, 218, 220, 238, 239, 369, 376, 417, 433, 434 Wilson, George, 418, 432 Winkler, Johann Heinrich, 200 Winter, Alison, 31, 51 Winterl, Jacob Joseph, 31, 121, 260, 261, 275–288, 313, 314, 318, 327, 369, 376 Wittgenstein, Ludwig, 243 Wöhler, Friedrich, 424 Wolff, Christian, x, 6, 76 Wollaston, William Hyde, 292, 293, 420 Wrede, Ernst Friedrich, 275 Wünsch, 99 Y Yelin, Julius Conrad von, 307–309, 312, 346 Z Zantedeschi, Francesco, 350 Zimmermann, E.A.W., 388–390

Hans Christian Ørsted and the Romantic Legacy in Science: Ideas, Disciplines, Practices (Boston Studies in the Philosophy of Science, Vol. 241) - PDF Free Download (2024)
Top Articles
3883 38th St, San Diego, CA 92105 - MLS 240019316 - Coldwell Banker
Beer Alcohol Content ABV% of Popular & Craft Beer Brands
Cremation Services | Mason Funeral Home serving Westfield, New York...
Brett Cooper Wikifeet
Rickrolling Link Generator
Guardians Of The Galaxy Showtimes Near Athol Cinemas 8
Mets Game Highlights
Northern Whooping Crane Festival highlights conservation and collaboration in Fort Smith, N.W.T. | CBC News
What Is Flipping Straights Ted Lasso
Quest Diagnostics Bradenton Blake - Employer Drug Testing Not Offered
Integrations | Information Technology
Sauce 423405
Rogers Breece Obituaries
Nutrislice White Bear Lake
5 high school boys cross country stars of the week: Sept. 13 edition
Elisabeth Fuchs, Conductor : Magazine : salzburg.info
Nail Shops Open Sunday Near Me
Estrella Satánica Emoji
Rick Harrison Daughter Ciana
Praxis für Psychotherapie und Coaching Rhein-Neckar
Ck3 Diplomatic Range
Xsammybearxox
Busted Newspaper Hampton County VA Mugshots
How Much Is Felipe Valls Worth
Orlando Magic Account Manager
Who are Yasmin Le Bon Parents? Meet Iradj Parvaneh and Patricia Parvaneh
Dominion Post Obituaries Morgantown
Prey For The Devil Showtimes Near Amc Ford City 14
Exploring Green-Wood Cemetery: New York Citys First Garden Cemetery | Prospect Park West Entrance,Brooklyn,11218,US | October 6, 2024
Odawa Hypixel
Satta King Peshawar
Where Is Gobblestone Castle
Eros Cherry Hill
3 Hour Radius From Me
Rek Funerals
Dl 646
Brett Cooper Wikifeet
Go Smiles Herndon Reviews
Top Chef Airer Nyt Crossword Clue
Mission Impossible 7 Showtimes Near Regal Willoughby Commons
Aces Login Palo Alto
Whats On Metv Now
Yakini Q Sj Photos
Ace Adventure Resort Discount Code 2023
Thoren Bradley Lpsg
Old Navy Student Discount Unidays
Albertville Memorial Funeral Home Obituaries
Oxford House Peoria Il
Evangeline Shrine Club Banquet Hall Photos
Fraction Button On Ti-84 Plus Ce
Creed 3 Showtimes Near Island 16 Cinema De Lux
Watch It Horror Thriller movies | Crystal panel
Latest Posts
Article information

Author: Twana Towne Ret

Last Updated:

Views: 5764

Rating: 4.3 / 5 (44 voted)

Reviews: 83% of readers found this page helpful

Author information

Name: Twana Towne Ret

Birthday: 1994-03-19

Address: Apt. 990 97439 Corwin Motorway, Port Eliseoburgh, NM 99144-2618

Phone: +5958753152963

Job: National Specialist

Hobby: Kayaking, Photography, Skydiving, Embroidery, Leather crafting, Orienteering, Cooking

Introduction: My name is Twana Towne Ret, I am a famous, talented, joyous, perfect, powerful, inquisitive, lovely person who loves writing and wants to share my knowledge and understanding with you.