The Significance of Indeterminacy: Perspectives from Asian and Continental Philosophy 2018019674, 9781138503106, 9781315145228

While indeterminacy is a recurrent theme in philosophy, less progress has been made in clarifying its significance for v

370 119 4MB

English Pages [403] Year 2018

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Significance of Indeterminacy: Perspectives from Asian and Continental Philosophy
 2018019674, 9781138503106, 9781315145228

Table of contents :
Cover
Title
Copyright
Contents
Acknowledgements
Preface
The Emerging Philosophical Recognition of the Significance of Indeterminacy
Part I The Significance of Indeterminacy in German Idealism
1 Overdeterminancy, Affirming Indeterminacy, and the Dearth of Ontological Astonishment
2 Determinacy, Indeterminacy, and Contingency in German Idealism
3 Free Thinking in Schelling’s Erlangen Lectures
4 Indeterminacy, Modality, Dialectics: Hegel on the Possibility Not to Be
Part II The Significance of Indeterminacy for Phenomenology, Natural Science, and Ethics
5 Determinable Indeterminacy: A Note on the Phenomenology of Horizons
6 Climate Science, Indeterminacy, and Food Security in Sub-Saharan Africa
7 Genetic Phenomenology and the Indeterminacy of Racism
8 Indeterminacy as Key to a Phenomenological Reinterpretation of Aristotle’s Intellectual Virtues
9 The Effability of the Normative
Part III The Significance of Indeterminacy for Hermeneutics and Aesthetics
10 Indeterminacy, Gadamer, and Jazz
11 Hermeneutic Priority and Phenomenological Indeterminacy of Questioning
12 Against the Darkness: Beauty and Indeterminacy in John Williams’s Stoner
13 Confidence Without Certainty
Part IV Asian Perspectives and Cosmological Concerns
14 Heidegger and Dōgen on the Ineffable
15 The Nietzschean Bodhisattva—Passionately Navigating Indeterminacy
16 Body and Intimate Caring in Confucian Ethics
17 Indeterminacy in Chinese Thought: Spontaneity and the Dao
18 Cosmological Questions
List of Contributors
Index

Citation preview

This topical and diverse collection of essays extends the critical and consequential problem of indeterminacy into both Continental and Comparative traditions. Creative yet rigorous, these essays enliven our sense of philosophy’s powers, defending the delicate ambiguity yet resonant force of philosophical claims as well as extending it to include traditions as varied as Buddhism and climate change policy. —Jason M. Wirth, Seattle University, USA

The Significance of Indeterminacy

While indeterminacy is a recurrent theme in philosophy, less progress has been made in clarifying its significance for various philosophical and interdisciplinary contexts. This collection brings together early-career and well-known philosophers—including Graham Priest, Trish Glazebrook, Steven Crowell, Robert Neville, Todd May, and William Desmond—to explore indeterminacy in greater detail. The volume is unique in that its essays demonstrate the positive significance of indeterminacy, insofar as indeterminacy opens up new fields of discourse and illuminates neglected aspects of various concepts and phenomena. The essays are organized thematically around ­indeterminacy’s impact on various areas of philosophy, including post-­ Kantian ­ idealism, phenomenology, ethics, hermeneutics, aesthetics, and East Asian philosophy. They also take an interdisciplinary approach by elaborating the conceptual connections between indeterminacy and literature, music, religion, and science. Robert H. Scott is an Assistant Professor of Philosophy at the University of North Georgia. His research focuses on phenomenology and environmental ethics, and in recently published work he develops a phenomenological theory of ecological responsibility. Dr. Scott currently serves as the President of the Georgia Philosophical Society. Gregory S. Moss is currently an Assistant Professor of Philosophy at Chinese ­University of Hong Kong. He specializes in post-Kantian German philosophy and has published in a variety of philosophical journals, such as Idealistic ­Studies, ­International Philosophical Quarterly, the Journal for the British Society for ­Phenomenology, Journal of Speculative Philosophy, and the Northern European Journal of Philosophy (forthcoming). Before completing his PhD on Hegel’s Logic of the Concept under Richard Winfield, he was a Fulbright Fellow with Markus Gabriel at the Rheinische Friedrich-Wilhelms-Universität Bonn. He is author of Ernst Cassirer and the Autonomy of Language and translator for Markus Gabriel’s Why the World Does Not Exist. His book Hegel’s Foundation Free Metaphysics: The Logic of Singularity is forthcoming with Routledge.

Routledge Studies in Contemporary Philosophy

Epistemic Rationality and Epistemic Normativity Patrick Bondy From Rules to Meanings New Essays on Inferentialism Edited by Ondřej Beran, Vojtěch Kolman, and Ladislav Koreň Toleration and Freedom from Harm Liberalism Reconceived Andrew Jason Cohen Voicing Dissent The Ethics and Epistemology of Making Disagreement Public Edited by Casey Rebecca Johnson New Directions in the Philosophy of Memory Edited by Kourken Michaelian, Dorothea Debus, and Denis Perrin A Pragmatic Approach to Libertarian Free Will John Lemos Consciousness and Physicalism A Defense of a Research Program Andreas Elpidorou and Guy Dove The Value and Limits of Academic Speech Philosophical, Political, and Legal Perspectives Edited by Donald Alexander Downs and Chris W. Surprenant The Significance of Indeterminacy Perspectives from Asian and Continental Philosophy Edited by Robert H. Scott and Gregory S. Moss For a full list of titles in this series, please visit www.routledge.com

The Significance of Indeterminacy Perspectives from Asian and Continental Philosophy Edited by Robert H. Scott and Gregory S. Moss

First published 2019 by Routledge 711 Third Avenue, New York, NY 10017 and by Routledge 2 Park Square, Milton Park, Abingdon, Oxon OX14 4RN Routledge is an imprint of the Taylor & Francis Group, an informa business © 2019 Taylor & Francis The right of the editors to be identified as the authors of the editorial material, and of the authors for their individual chapters, has been asserted in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988. All rights reserved. No part of this book may be reprinted or reproduced or utilised in any form or by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying and recording, or in any information storage or retrieval system, without permission in writing from the publishers. Trademark notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe. Library of Congress Cataloging-in-Publication Data Names: Scott, Robert H. (Assistant Professor of Philosophy), editor. Title: The significance of indeterminacy : perspectives from Asian and Continental philosophy / edited by Robert H. Scott and Gregory S. Moss. Description: 1 [edition]. | New York : Taylor & Francis, 2018. | Series: Routledge studies in contemporary philosophy ; 110 | Includes bibliographical references and index. Identifiers: LCCN 2018019674 | ISBN 9781138503106 (hardback) Subjects: LCSH: Determinism (Philosophy) | Free will and determinism. | Certainty. | Continental philosophy. Classification: LCC B105.D47 S525 2018 | DDC 123—dc23 LC record available at https://lccn.loc.gov/2018019674 ISBN: 978-1-138-50310-6 (hbk) ISBN: 978-1-315-14522-8 (ebk) Typeset in Sabon by Apex CoVantage, LLC

Contents

Acknowledgements Preface

x xi

ROBERT H. SCOTT AND GREGORY S. MOSS



The Emerging Philosophical Recognition of the Significance of Indeterminacy

1

GREGORY S. MOSS, C.U.H.K.

PART I

The Significance of Indeterminacy in German Idealism

49

1 Overdeterminancy, Affirming Indeterminacy, and the Dearth of Ontological Astonishment 

51

WILLIAM DESMOND

2 Determinacy, Indeterminacy, and Contingency in German Idealism

67

G. ANTHONY BRUNO

3 Free Thinking in Schelling’s Erlangen Lectures

84

GREGORY S. MOSS

4 Indeterminacy, Modality, Dialectics: Hegel on the Possibility Not to Be104 NAHUM BROWN

viii Contents PART II

The Significance of Indeterminacy for Phenomenology, Natural Science, and Ethics

125

  5 Determinable Indeterminacy: A Note on the Phenomenology of Horizons

127

STEVEN G. CROWELL

  6 Climate Science, Indeterminacy, and Food Security in SubSaharan Africa

148

TRISH GLAZEBROOK AND MICHAEL GOLDSBY

  7 Genetic Phenomenology and the Indeterminacy of Racism

168

JANET DONOHOE

  8 Indeterminacy as Key to a Phenomenological Reinterpretation of Aristotle’s Intellectual Virtues

182

ROBERT H. SCOTT

  9 The Effability of the Normative

201

TODD MAY

PART III

The Significance of Indeterminacy for Hermeneutics and Aesthetics

213

10 Indeterminacy, Gadamer, and Jazz

215

BRUCE E. BENSON

11 Hermeneutic Priority and Phenomenological Indeterminacy of Questioning

228

NATHAN ERIC DICKMAN

12 Against the Darkness: Beauty and Indeterminacy in John Williams’s Stoner247 PHILLIP E. MITCHELL

13 Confidence Without Certainty J. AARON SIMMONS

260

Contents  ix PART IV

Asian Perspectives and Cosmological Concerns

277

14 Heidegger and Dōgen on the Ineffable

279

FILIPPO CASATI AND GRAHAM PRIEST

15 The Nietzschean Bodhisattva—Passionately Navigating Indeterminacy309 GEORGE WRISLEY

16 Body and Intimate Caring in Confucian Ethics

330

QINGJIE JAMES WANG

17 Indeterminacy in Chinese Thought: Spontaneity and the Dao

342

ROBERT CUMMINGS NEVILLE

18 Cosmological Questions

357

RICKI BLISS AND FILIPPO CASATI

List of Contributors Index

375 380

Acknowledgements

Many teachers and interlocutors have contributed to the development of the ideas that have come together in this volume. First, the editors would like to thank all of the chapter contributors for their excellent and tireless work on their chapters. Second, we would like to recognize and thank Dr. O. Bradley Bassler (University of Georgia) in whose 2008 seminar on Metaphysics, which both editors took part in, the seed for this work was planted and began to grow. Robert H. Scott would like to thank the University of North Georgia for providing generous support for his work on this volume during the summer of 2017 through the UNG “Presidential Summer Incentive” grant program. Both editors would like to thank the editorial staff of the Studies in Contemporary Philosophy series at Routledge and, in particular, Andrew Weckenmann, for his patience and support throughout the editorial process.

Preface

With the diversification of philosophy, and the dismantling of stark divides in philosophical methodology in the West, the character of philosophy appears more indeterminate than ever—and demands fresh investigations not only into the character of philosophy, but also the concept of indeterminacy itself. The over-arching aim of this collection is to bring into focus the prominence and significance of indeterminacy as a common thread in recent Asian philosophy, continental thought, and other philosophical approaches. The theme of indeterminacy can be traced throughout the history of both Western and Asian philosophy. Among the pre-Socratics, Anaximander stands out as recognizing (though not fully clarifying) its significance in his famous formulation of the first principle of philosophy as the apeiron—the indeterminate. In modern philosophy, indeterminacy appears time and again as a recurrent theme in post-Kantian idealism, phenomenology, and other areas of continental philosophy. This volume shines a spotlight on the way indeterminacy arises as an important theme for relatively neglected thinkers in the Western tradition, such as F.W.J. Schelling, and it offers fresh perspectives on the significance of indeterminacy for well-read thinkers such as Husserl and Hegel. What is more, this volume includes chapters that bring out the presence and importance of indeterminacy in the various schools of Chinese and Japanese philosophy (among others), such as in various forms of Daoist thinking and the Kyoto School, which cannot be underestimated. By bringing these schools of thought into dialogue with each other, we hope that the volume will enrich the thinking of all traditions, East, West, and beyond. Of the many books that have appeared on indeterminacy,1 only a few edited volumes on indeterminacy have appeared.2 This volume fills a significant lacuna in the English-speaking literature by exploring indeterminacy from various angles within the Continental and comparative philosophical traditions. Although the volume focuses on what has traditionally been called “Continental philosophy” many of our authors are also trained in the Analytic tradition, such as Graham Priest, and the editors have embraced a deeply inter-disciplinary approach that includes the perspectives of eighteen diverse prominent and emerging scholars on the issue, including five authors

xii Preface who focus on the importance of indeterminacy for Asian philosophy. Even in monograph form, indeterminacy appears to be a significantly neglected theme in Asian thought in the English-speaking world.3 Whereas many of the texts that have been published on the topic treat indeterminacy as an important theme for a particular subject such as economic theory, semantic theory, music theory, or theories of action, this volume takes the reverse approach by starting with indeterminacy as the primary theme from which discussion of its significance for various spheres of philosophical inquiry, including philosophy of art, ethics, hermeneutics, and Asian philosophy, may proceed. In our view, the globalization and diversification of philosophy demands a serious comparative engagement with the significance of indeterminacy as it arises as a theme in the East, West, and beyond. Thus, while the facticity of indeterminacy has been recognized in various contexts and areas, less progress has been made in clarifying its significance for the various contexts and disciplines in which it arises. Moreover, where attempts have been made to clarify its significance in the Western tradition, the tendency has been to frame the indeterminate in predominately negative terms. An important aim of this volume is to bring into focus ways in which indeterminacy carries and has carried positive significance for philosophical thinking. To this end, most (though not all) of the chapters address the positive significance of indeterminacy. Some of the chapters show the role of indeterminacy in drawing links to ethics, and some address its role in making various inter-disciplinary links. In regard to the latter, various chapters explore the significance of indeterminacy in relation to science (especially chapter 6), religion (e.g., chapters 3 and 13–18), fiction (chapter 12), and music (chapter 10). We hope these pages will provide the reader with an impetus and a starting point for further discussion and investigation into the role that indeterminacy plays in the development of a wide-ranging global and inter-disciplinary discourse. Naturally, this volume purports to be neither an exhaustive study of the various ways indeterminacy (and especially its positive significance) have appeared in the history of philosophy, nor a complete study of all of the possible directions of inquiry that might be of relevance to the issue at hand. Rather, our hope is that The Significance of Indeterminacy: Perspectives from Asian and Continental Philosophy will stimulate the thinking of a wide and diverse group of philosophers and students alike on the theme of indeterminacy, a theme which will certainly be of great import for the future of philosophy and its self-understanding in the twenty-first century. Robert H. Scott and Gregory S. Moss

Notes 1  In regard to the existing literature, in Continental philosophy the theme of indeterminacy in Husserl and those influenced by him has been directly addressed by Weiss (2008). There are a few books on the significance of indeterminacy for linguistics and philosophy of language: Friedrich (1996), Vogeleer et al. (1999), McCarthy (2002),

Preface  xiii Aarts (2007), Librar (2013), Suess (2014), and Gaudet (2006). The following texts treat the issue of indeterminacy insofar as it relates to economic theory: Nishimura (2012) and Varoufakis (2015). There are a number of texts that investigate indeterminacy in respect to agency, causality, action, and political and legal philosophy: Eisenberg (1992), Gregg (2003), Hardin (2005), Deutschmann (2013), Schatzky (2012), and Mueller (2014). The following texts primarily concern the relation between pragmatism and indeterminacy: Rosenthal (2000) and Cooke (2007). There are also a number of books that address the issue of indeterminacy in philosophy of mind and aesthetics: Schoffman (1990); Perloff (1999), Danvers (2006), Wurth (2009), and Pavón-Cuéllar and Parker (2013). Other texts investigate the relation between indeterminacy and the sciences, or have a special focus on indeterminacy in analytic philosophy: Bohman (1993) and Ciprut (2009). The following texts are primarily concerned with indeterminacy in respect to epistemological and metaphysical concerns: David (2011), Martine (1992), Parsons (2000), Correia and Iacona (2013), and Bassler (2015). Other edited volumes on the theme of indeterminacy include Vogeleer et al. 2   (1999), Ciprut (2009), and Mueller (2014). 3  An exception is Danvers (2006).

Works Cited Aarts, Bas. 2007. Syntactic Gradience: The Nature of Grammatical Indeterminacy. Oxford: Oxford Linguistics. Bassler, O. Bradley. 2015. The Long Shadow of the Parafinite: Three Scenes from the Prehistory of a Concept. Chestnut Hill, MA: Docent Press. Bohman, James. 1993. New Philosophy of Social Science: Problems of Indeterminacy. Cambridge, MA: MIT Press. Ciprut, Jose V. ed. 2009. Indeterminacy: The Mapped, the Navigable, and the Uncharted. Cambridge, MA: MIT Press. Cooke, Elizabeth. 2007. Peirce’s Pragmatic Theory of Inquiry: Fallibilism and Indeterminacy. London: Bloomsbury Studies in American Philosophy. Correia, Fabrice and Andrea Iacona. eds. 2013. Around the Tree: Semantic and Metaphysical Issues Concerning Branching and the Open Future: 361. Dordrecht: Springer, Synthese Library. Danvers, John. 2006. Picturing Mind: Paradox, Indeterminacy and Consciousness in Art & Poetry. In Consciousness, Literature, and the Arts 3. Amsterdam: Rodopi BV. David, A. P. 2011. Plato’s New Measure: The ‘Indeterminate Dyad’. Createspace Independent Publishing Platform. Deutschmann, Moritz. 2013. Free will, Indeterminacy, and Self-determination. Munich: GRIN, Verlag. Eisenberg, John A. 1992. The Limits of Reason: Indeterminacy in Law, Education, and Morality. New York, NY: Routledge. Friedrich, Paul. 1996. The Language Parallax: Linguistic Relativism and Poetic Indeterminacy. Austin, TX: University of Texas Press, Texas Linguistics Series. Gaudet, Eve. 2006. Quine on Meaning: The Indeterminacy of Translation. London: Bloomsbury Studies in American Philosophy. Gregg, Benjamin. 2003. Coping in Politics with Indeterminate Norms: A Theory of Enlightened Localism. Series in Radical Social and Political Theory: Contemporary Issues. New York, NY: SUNY Press.

xiv Preface Hardin, Russell. 2005. Indeterminacy and Society. Princeton: Princeton University Press. Librar, Adnan. 2013. Residual Indeterminacy or Optionality at Core Syntax. Riga: Lambert Academic Publishing. Martine, Brian John. 1992. Indeterminacy and Intelligibility. Series in Systematic Philosophy. New York, NY: SUNY Press. McCarthy, Timothy. 2002. Radical Interpretation and Indeterminacy. Oxford: Oxford University Press. Mueller, Thomas. ed. 2014. Nuel Belnap on Indeterminism and Free Action. Dordrecht: Springer, Outstanding Contributions to Logic. Nishimura, Kazuo. 2012. Nonlinear Dynamics in Equilibrium Models: Chaos, Cycles and Indeterminacy, edited by John Stachurski, Alain Venditti and Makoto Yano. Dordrecht: Springer. Parsons, Terence. 2000. Indeterminate Identity: Metaphysics and Semantics. Gloucestershire: Clarendon Press. Pavón-Cuéllar, David and Ian Parker. 2013. Lacan, Discourse, Event: New Psychoanalytic Approaches to Textual Indeterminacy. New York, NY: Routledge. Perloff, Marjorie. 1999. The Poetics of Indeterminacy: Rimbaud to Cage. Evanston, IL: Northwestern Universtiy Press, Avant-Garde & Modernism Studies series. Rosenthal, Sandra B. 2000. Time, Continuity and Indeterminacy: A Pragmatic Engagement with Contemporary Perspectives. New York, NY: SUNY Press. Schatzky, Theodor R. 2012. The Timespace of Human Activity: On Performance, Society, and History as Indeterminate Teleological Events. Lexington, KY: Lexington Books. Schoffman, Nachum. 1990. From Chords to Simultaneities: Chordal Indeterminacy and the Failure of Serialism. Contributions to the Study of Music and Dance. Santa Barbara, CA: Praeger. Suess, Moti. 2014. A Semantic Account of Vagueness: Vagueness as Indeterminacy. Riga: Lambert Academic Publishing. Varoufakis, Yanis. 2015. Economic Indeterminacy: A Personal Encounter with the Economists’ Peculiar Nemesis. New York, NY: Routledge, Frontiers of Political Economy. Vogeleer, Svetlana, Walter De Mulder, and Ilse Depraetere. eds. 1998. Tense and Aspect: The Contextual Processing of Semantic Indeterminacy. Belgian Journal of Linguistics 12, Amsterdam: John Benjamins Publishing Company. Weiss, Gail. 2008. Refiguring the Ordinary. Bloomington: Indiana University Press. Wurth, Kiene. 2009. Muscially Sublime: Indeterminacy, Infinity, Irresolvability. Bronx, NY: Fordham University Press.

The Emerging Philosophical Recognition of the Significance of Indeterminacy1 Gregory S. Moss, C.U.H.K.

The Problem of Indeterminacy in the History of Philosophy The Significance of Determinacy Although Anaximander imbues ἄπειρον with positive significance (610–540 BC) as the principle of all things,2 the overwhelming influence of the peripatetic philosophy in the Western tradition relegated the ἄπειρον to a problematic and negative position. Indeed, in the peripatetic tradition we discover one important source for the perennial tendency to privilege determinacy over indeterminacy in the Western tradition. Determinacy, understood in the sense of the Latin determinatio signifies “limit” or “end.” In Greek, we can trace back the determinate to the Πέρας (limit), and the indeterminate to its negation, ἄπειρον (the indeterminate or unlimited), in which the alpha-­ privative signifies that indeterminacy is a lack of determinacy. Whether wisdom be identified with the principles of Being or knowing, even from the Greek language itself we find that it is not Πέρας that is defined by negation, but ἄπειρον. Aristotle, as a friend of wisdom—φιλοσοφία—is not merely concerned with knowing particular kinds of wisdom, but with the principles upon which all wisdom depend. Since all wisdom has Being (εἶναι) as its object, the first principles of wisdom are the first principles of Being. To state the “what it is to be” (τὸ τί ἦν εἶναι) of things requires discovering what the thing is in its own right (καθ᾽ αὑτὸ).3 What exists in its own right does not exist relative to or in virtue of something else, and is thereby independent. These independent principles are not infinite (οὐκ ἄπειρα), and are responsible (αἰτία) for what is.4 Indeed, Aristotle’s identification of the principle of Being with what is limited and independent does not begin with him. Plato before him identifies each Form (or εἶδος) as αὐτὸ καθ᾽ αὑτὸ: itself by itself.5 For Plato and Aristotle, Being is identified with Form, and Form is an independent thing, which exists in virtue of itself (καθ᾽ αὑτὸ). Knowing the being of wisdom requires knowing the being of philosophical knowing. Thus, it falls to the philosopher to know what it is to be a friend of wisdom. Following this line of inquiry, philosophy heeds the oracle

2  Gregory S. Moss, C.U.H.K. at Delphi to know oneself, γνῶθι σεαυτόv. Philosophy is free, or autonomous, in the sense that no other discipline can determine what philosophy is or ought to be. Philosophy must determine itself. In his poem, Parmenides famously identifies Being as finite or determinate, ἐν πείρασι. On the one hand, Being is that which is determinate insofar as it does not exceed its own limit. Being cannot be Non-being. On the other hand, Non-being does exceed its own limit: whenever one thinks of Nonbeing, one thinks some being. Because Non-being slips into its opposite, it fails to remain within its limit: it is ἄπειρον: that which exceeds its own limit. Non-being has the character of what Heraclitus affirms as παλίντροπός6— it turns back on itself. For Parmenides, and much of Greek thought, the determinate is one: it is just Being. The indeterminacy of Non-being is dissolution into many: it is and it is not. Plato adopts this distinction in the Phaedo: The things that are (ta onta), such as the Form of Equality never becomes unequal. Forms stay themselves and do not become what they are not: they are not indeterminate, for they do not exceed their own limit. Each is indivisibly one. In the Philebus Plato identifies what is indeterminate with Becoming, namely what always introduces a comparison, in particular that of more and less,7 in contrast with fixed quantities and forms such as Equality, which do not admit of more and less and do not change.8 In his own way, Aristotle follows this path too by arguing that Being is governed by the principle of non-contradiction. From Parmenides through Plato and Aristotle, Being appears again and again as determinate: it is that which is independent or by itself, and does not exceed its own limit, while the indeterminate is what exceeds its own limit, and is inseparable from plurality and comparative determinations. Given that for Aristotle Being is that which ultimately has a limit [πέρας], and its limit constitutes that point at which it has come to its end, each thing is complete only insofar as it reaches its end. Since ἄπειρον is that which always has something outside it, it is that which never comes to an end. Thus, the indeterminate is always incomplete. For what has reached its end, nothing integral to its own being is left outstanding: it is complete and whole. Thus, for Aristotle the infinite does not exist as an independent thing, for each independent thing is complete and a whole. Rather, like Aristotle’s concept of matter, the infinite only exists potentially. To take a simple example, although the line is a continuous quantity that does not consist of an infinite number of points, the line has the potential to be divided without end. The line has ἄπειρον potentially, and when that potential is actualized, the infinite series of divisions is only present as something incomplete, as a series in which one more division can always be taken.9 Just as Being is primarily understood as something determinate, so too is the Ancient inquiry into and science of Being determinate. In order to know Being, knowing must conform to the structure of Being. Accordingly, Parmenides identifies thought (νοῦς) with Being. In Plato’s Meno and Phaedo, learning is recollection, and the object of recollection is the Form, not

Emerging Recognition of the Significance of Indeterminacy  3 the world of Becoming that is accessible by means of the senses. Moreover, in Republic it is by Nous that one knows the determinate Forms. Even if philosophy never arrives at science, as is at least the case in Plato’s early dialogues, the possibility of philosophical inquiry and indictment of contradictions as necessary falsehoods points to a strong commitment to the determinate character of knowing, even if humans fail to possess it. Finally, Aristotle’s syllogistic logic is designed to deduce the per se attributes of things from the middle term, which is the logical correlate of the Being of the thing as it exists in its own right (καθ᾽ αὑτὸ). To know is to grasp a limit; there is no proper science of the indeterminate. Perhaps there is no better illustration of the priority of the πέρας in peripatetic thought than in Aristotle’s concept of ἐντελέχεια: being-at-work staying itself. Nous is one of Aristotle’s primary examples of ἐντελέχεια. It is an activity in which the end (purpose) is not distinct from the means. As Aristotle states: “one sees and at the same time in a state of having seen, [. . .] one thinks contemplatively, and is at the same time in a state of having thought contemplatively.”10 This is different from house building, for example, where in the act of building the house one is not in the state of having built the house. Unlike in house building, when one achieves the end or purpose of Nous, namely when one knows the first principle, the means by which one knows does not cease to be. Because the end is the means, the means continues to be with the achievement of the end. Nous is an end-in-itself. Aristotle is clear that what exists in virtue of itself (καθ᾽αὑτὸ) is self-sufficient and complete. For Aristotle, what has nothing outside it is complete, and “nothing is complete which has no end (tέλος), and the end is a limit.”11 In order to be complete something must have a limit (πέρας) within itself. In the case of ἐντελέχεια, the activity is complete because the end or the limit is identical with the means and is not outside it: there is no instance where the means is present and the end is not. In Nous, the limit is present at every instance of the activity: it is an act in which the limit is one (ἐν) with the purposive activitiy (τελέχεια). For Aristotle, philosophical knowing of first principles is a knowing that is not ἄπειρον, for it never oversteps its own limit; rather it maintains itself at its limit (πέρας). It is the (πέρας) or determinate that is perfect and complete in knowing and Being. Aristotle famously argued that philosophy is the freest of the sciences.12 A person is free who is an end in himself, and philosophy is the only science that seeks knowledge as an end in itself. Thus, philosophy is free. In the practice of Nous, the philosopher practices a kind of knowing in which means and ends are identified. Accordingly, Nous is free knowing. In sum, autonomous knowing means (i) knowing what is autonomos or καθ᾽αὑτὸ, and (ii) knowing it in an autonomous way: free knowing in Aristotle may appear ironic to the modern reader, for it is bound and determinate. Free knowing never exceeds its own limits. Immanuel Kant’s Copernican Revolution upends Being as the privileged object of knowing. Rather than directly inquire into Being, it falls to the

4  Gregory S. Moss, C.U.H.K. philosopher to inquire into the capacity of reason to know its object. Thus, the primary object of philosophical inquiry is no longer Being per se, but knowing. Although the Copernican Revolution upends Being as the privileged object of knowing, Kant’s construal of philosophy as the knowing of knowing nonetheless follows in the spirit of Aristotle’s description of philosophy as free science, for it falls to philosophy to determine the limits of its own knowing. What is more, Kant appears to continue to privilege, in his own way, determinacy over indeterminacy in his construal of the structure of knowledge. In this same spirit of autonomous self-inquiry, Kant, in his Critique of Pure Reason, engages in a philosophical reflection on the limits of philosophical knowing. As autonomous self-inquiry, philosophy is charged with charting not only the characteristic determination(s) of autonomous and non-autonomous knowing, but also the very character of determinacy itself. In his discussion of the Transcendental Ideal, Kant formulates one of two principles of determination: Grundsatz der Bestimmbarkeit, namely the principle of determinability: “every concept, in regard to what is not contained in it, is indeterminate [unbestimmt], and stands under the principle of determinability: that of every two contradictorily opposed predicates only one can apply to it, which rests on the principle of contradiction.” This principle of determinability “has nothing in view but the logical form of cognition [Erkenntnis].”13 Following Kant’s lead, in respect to one pair of contradictory opposites, such as F and not-F, every concept (in regard to predicates that are not already contained in the concept) can be made determinate by applying only one side of the pair of opposition to the concept. If both predicates were applied to the same concept, then the concept would admit both F and not-F, which would be a contradiction. This principle is formal, since it abstracts from the content of each predicate. Accordingly, the principle of determinability stands under the law of non-contradiction and thereby restricts all predications of a concept to affirm at most only one side of any pair of opposites. By contrast, the concept remains indeterminate insofar as the concept is neither F nor not-F. To determine the concept, one side of the opposition must be negated. It would appear that the concept of negation is in some sense derivative,14 for it requires that some affirmative content or predicate be given in advance of the negation. In his table of categories, ‘negation’ appears as the second category of three under Quality: reality, negation, and limitation.15 The third category is a synthesis of the first two: limitation is reality combined with negation.16 Indeed, as a category, it is not derivative, but underived. For a concept to become determinate, the possible predicates that could apply to it must be limited. Accordingly, determination requires both the concepts of reality (in this case, some affirmative conceptual content) and negation. Following the tradition, Kant ascribes the law of non-contradiction to the form of logical knowing, such that only determinable concepts can count as abiding by the

Emerging Recognition of the Significance of Indeterminacy  5 form of logical knowing. In this vein, although contradictions are necessarily false, consistent propositions are not necessarily true, but are candidates for truth, for the concept could still fail to correspond with its object. Following the tradition, Kant restricts the form of logical cognition to determinate knowing, according to which determinacy has been privileged over indeterminacy as a necessary condition for entering into the domain of truth. In his Science of Logic, Hegel further develops the relationships between reality, negation, limitation, and determinacy. Hegel generally describes determinate being as “being with limitation or negation.”17 “Determinateness is a negation posited as affirmative.”18 Moreover, determinate being is “the sphere of difference, of dualism, the field of finitude.”19 One of Hegel’s many insights is that every determinate concept already contains both reality and negation. Each concept has its own conceptual content—this is its reality. This reality is determinate if it excludes or negates what is other to it. Accordingly, what is determinate negates what is other, and thereby maintains its own independence apart from the other. Determinacy seems to be constituted by a structure of something and other, each of which is its own reality and negates what is distinct from it. Accordingly, whenever it is true to say of something that it is “something” or “other,” one is operating within the sphere of determinacy. This dualism can be described in different ways as the dualism of being and nothing, reality and negation, or something and other. Since each something is in relation to an other, determinate being appears to be relative, rather than absolute.20 This is reflected in the etymology of the term Dasein, which Hegel defines as determinate being: “Dasein ist bestimmtes Sein.”21 “Being-there is determinate being,” or as Giovanni translates it: “existence is determinate being.”22 The etymology of Dasein is being (Sein) in a certain place (da).23 To be determinate is not just to be, but to be in a position. This position, of course, is not necessarily spatial, for something can be determinate and be a non-spatial concept, e.g., a number. Even before one determines a concept by negating not-F and affirming F, F and not-F each already contain negation in order to be determinate predicates and limit one another in the first place. Determinate being is “illuminated darkness” and “darkened light.”24 Indeed, Hegel seems right to follow Spinoza’s famous dictum that omnis determinatio est negatio. The Problem of Indeterminacy Prima facia, it might appear that indeterminacy ought not pose any special philosophical problem. Indeterminacy is the negation of the Latin determinatus, which signifies “limit.” Accordingly, indeterminacy is the absence of determinacy, the absence of limit. As an absence of determinacy, it has a negative significance, for it is the non-being of determinacy. If determinacy is established as a condition for the possibility of truth, indeterminacy as such is not possibly true. Accordingly, the negative significance of indeterminacy

6  Gregory S. Moss, C.U.H.K. is at least twofold: it is the non-being of determinacy, and cannot be true. This twofold negativity of indeterminacy is reflected in Aristotle’s insightful claim that the term “falsehood” is another way of saying “non-being.” To put this in the strongest possible terms, we might say that indeterminacy falls afoul of metaphysics and epistemology: it is a false nothingness. Following the German etymology, since determinacy is being-in-position, indeterminacy must be that which is non-positional. Although indeterminacy may appear to be philosophically unproblematic, the history of philosophy has decided otherwise. If we follow the concept of non-positionality, we discover a paradox lurking beneath. The paradox is nicely illustrated by a famous story from the Daoist text, Zhuangzi: The emperor of the southern sea was called Swoosh. The emperor of the northern sea was called Oblivion. The emperor of the middle was called Chaos. Swoosh and Oblivion would sometimes meet in the territory of Chaos, who always attended to them quite well. They decided to repay Chaos for his virtue. “All men have seven holes in them, by means of which they see, hear, eat, and breathe,” they said. “But this one alone has none. Let’s drill him some.” So each day they drilled another hole. After seven days, Chaos was dead.25 Chaos is indeterminate. Swoosh and Oblivion attempt to give him determinacy by drilling holes into him. By doing so, Chaos receives form and becomes determinate. As a result, Chaos dies.26 If the indeterminate is the negation of determinacy, then it must be opposed to positional being. But if indeterminacy is placed in opposition to positional being, it must also have a position. Thus, a contradiction follows: the non-positional would be positional, or what is the same, the indeterminate would not be indeterminate, but determinate. Hegel explicitly acknowledges the same contradiction, but in less colorful terms: But indeterminateness is a determinateness, because it is supposed to stand opposed to the determinate. But the enunciation of what it is, itself sublates what it is supposed to be;27 Indeterminacy poses a significant philosophical problem: every time one attempts to conceive of the indeterminate, one places indeterminacy into an object of determinate knowing. As long as indeterminacy remains separate from positional knowing [or positional being], it cannot remain indeterminate. But this paradox illuminates a further feature of indetermination: since it cannot be opposed to determinate being without becoming determinate, it can only remain indeterminate by no longer remaining opposed to it. Thus, what is non-positional can only remain non-positional by occupying every position. Put in more formal terms, if indeterminacy is neither F nor

Emerging Recognition of the Significance of Indeterminacy  7 not-F (where ‘F’ is now taken to be a determination), it must also be both F and not-F. In virtue of excluding all possible determinations it must include all possible determinations. Indeed—the non-positionality of the indeterminate appears to be the conceptual source (ἀρχή) of the Ancient insight that the indeterminate exceeds its own limits. Because the indeterminate runs through all determinations, every determinate field of inquiry appears to be thoroughly saturated with indeterminacy. In order to acknowledge the ubiquity of indeterminacy, The Significance of Indeterminacy does not restrict the investigation of indeterminacy to any particular subject matter. Rather, every field of determinacy is run through with indeterminacy, and invites scholars to uncover the indeterminacy therein. Whether it is aesthetics, logic, or physics, indeterminacy is there (Da). Indeed, it is the very problem of determining the position of indeterminacy that invites an inter-disciplinary method of inquiry. Cognates of Indeterminacy Although the main theme of the volume is indeterminacy, a number of contributors focus on cognates of indeterminacy, such as indefinability, ineffability, inconceivability, unknowability, and uncertainty. Although it may at first appear that these concepts are separable from indeterminacy, if we follow the phenomenon of non-positionality, we discover an intimate connection between indeterminacy and these various concepts. Since indeterminacy is the negation of determinacy, and determinate being is a finite characterization, the indeterminate is not-finite, or infinite. Of course, there are many senses of infinitude, such as quantitative infinity, the Medieval concept of perfection (lacking nothing), and indeterminacy. The indeterminate appears to straddle all three. Because the indeterminate is without character, it has no-determination, and is thereby indistinguishable from all possible characteristics. For if it were opposed to any determinations, it would be determinate. Thus, the indeterminate is just as much an absence of all determinations as it is constituted by the presence of all determinations. Nahum Brown, in Indeterminacy, Modality, Dialectics: Hegel on the Possibility Not to Be not only works out a Hegelian conception of the possibility not to be, but also offer an insight into Hegel’s concept of the indeterminate totality with the emphasis on possibility: the whole is the totality of possibility: both A and not-A. Indeterminacy can just as much be linked to the positive sense of the infinite as the perfect, or that from which nothing is lacking. The indeterminate also appears to be connected to quantitative infinity. This can be most elucidated from its apparent inconceivability. If what can be known by means of concepts follows the principle of non-contradiction and excluded middle, one can only know that which is in a position—­ constrained to one side of a set of contradictory opposites. Yet, indeterminacy is constituted as both ‘neither F nor not-F’, and ‘both F and not-F.’ Thus, indeterminacy appears to elude conceptual knowing. If knowing is

8  Gregory S. Moss, C.U.H.K. constrained to discursive knowing, then the indeterminate may be further characterized as unknowable. For the quantitative infinite, at every position p in the order of succession, there is always another position p* that lies beyond p. Likewise, every attempt to think the indeterminate leads only to another form of determinacy, ad infinitum. The indeterminate is always a step beyond every position that is ascribed to it. Conceiving the indeterminate generates an infinite regress, a quantitative infinity, in which the indeterminate is eternally anticipated, though never conceived. Knowing becomes an infinite striving without end. Because the indeterminate appears to be a step beyond every position, every conception of the indeterminate is incomplete. But it is not only the conception of the indeterminate, but perhaps the indeterminate itself that also connotes incompleteness. Insofar as the indefinite lacks every position— it is what is non-positional, it is the exact opposite of perfection: it is fundamentally incomplete. On the one hand, in lacking positionality, the indeterminate as such is neither F nor not-F, and must be absolutely incomplete. On the other hand, whatever is relatively indeterminate is incomplete in some respect, rather than in every respect, as Kant indicates when he designates a concept as indeterminate insofar as it is determinable in respect to one of two opposed predicates that are not contained in the concept. The tendency of the indeterminate to overstep its own limit is reflected in the connotations of both completeness and incompleteness. What goes by the name of Gödel’s incompleteness theorem engenders a kind of incompleteness in the system of Arithmetic quite different from the sense of an infinite remainder. Put shortly, the kind of incompleteness at work here is un-decidability. Gödel showed that there are statements in Language F that can neither be proved nor disproved. In regard to provability, such statements remain un-decidable. The indeterminate is also uncertain. Certainty and uncertainty are psychological concepts. Insofar as we fail to conceive what the indeterminate is, we are not certain what the indeterminate is. Heisenberg’s Uncertainly Principle reflects a specialized sense of the indeterminate: we cannot simultaneously establish the position and the velocity of an electron. This principle raises questions about how our very interaction in the world as thinkers and scientists can generate indeterminacy. The establishment of the one entails uncertainty regarding the other. As indeterminate, the indeterminate evades our understanding: it is what cannot be pinned down. The indeterminacy is what does not settle; what settles takes a place, and indeterminacy never takes a place. The uncertainty of indeterminacy is linked with this pervasive evasiveness as it affects our psychological dispositions. We cannot be assured about the indeterminate. At best, perhaps we can be confident about its very evasiveness—its capacity to leave us unsettled. The indeterminate, as the placeless, does not appear to give us any place to dwell or inhabit. In this sense, the indeterminate also seems to connote, to borrow a phrase from Merleau-Ponty, a kind of wild being. In its absolute form,

Emerging Recognition of the Significance of Indeterminacy  9 it appears as pure Chaos (Χάος), the primordial god of Greek mythology. Chaos is a wild place where no dwelling is possible; it connotes a yawning gap or abyss. Many mythologies have the figure of the Chaoskampf: the struggle against chaos. By bringing order to chaos, chaos does not disappear—rather it appears to permeate every ordered whole. Whether we consider definition in the traditional sense of a genus in addition to the primary difference, definition by intension or extension, ostensive definitions, or functional definitions, in every case some positional specification is always involved. Differentiating humans from oxen requires importing the διαφορά, the differentia. Likewise, in order to define by ostension or extension more generally, one must differentiate the objects or relations in order to refer to them as this or that. Whatever can be differentiated has position vis-à-vis another. Indeterminacy is never “this” or “that”; it transcends all such positions, whether they be universal or particular. Thus, indeterminacy appears to be indefinable. Regarding meaningful speech, in book Г of his Metaphysics, Aristotle argues that one can only speak meaningfully if one’s terms do not mean everything.28 Only by excluding some meanings can a word become meaningful. To communicate one’s thoughts about Heidegger one must be able to distinguish the meaning of the term “Heidegger” from other possible referents. If meaningful speech and conversation requires that each of our words does not mean infinitely many things, then it is unclear whether meaningful speech about the indeterminate is possible. For it would appear that indeterminacy either means nothing in particular or is indistinguishable from all possible meanings. Thus, indeterminacy has been rightly linked with the ineffable and unnamable. So another paradox arises: indeterminacy seems to leave us without a word to say about it, and yet there is a whole volume before you. Although we might insist that indeterminacy means “non-positional,” this meaning immediately betrays a lack of sense, and thrusts us back out into the ἄπειρον. The Problem of Onto-theology

The ubiquity of indeterminacy indicates that indeterminacy cannot be limited. If determinacy is relative being, and indeterminacy is the absence of determinacy, then indeterminacy is the absence of relativity, or what is better: absolute. Being-in-position is always relative vis-à-vis another. But the impossibility of setting indeterminacy to one side of an opposition indicates that indeterminacy cannot be relative to another, but is unrelated to any others outside of it: it appears universal and all-encompassing. In a phrase of the young Idealists in Jena, the indeterminate appears to be the Ἓν καὶ Πᾶν, the one and all, for nothing lies beyond it, and it runs through all determinations. Unlike relative determinations, the Absolute is universal and allencompassing. Accordingly, the Absolute is not in position vis-à-vis another. Hence, the Absolute also appears to be non-relative or by itself in the same way the indeterminate is by itself.

10  Gregory S. Moss, C.U.H.K. As we formulate the fundamental problem of indeterminacy, we might first re-formulate the opposition of indeterminacy with determinacy: while the former appears to be by itself, the latter is in relation. What is by itself is independent and autonomous. Accordingly, an important locus of the investigation into indeterminacy involves the concepts of independence and autonomy. We should remember that Ancient Greek philosophy identified determinate existence with the property of independence expressed in the phrase auto kath auto, itself by itself. As philosophy comes into the 19th and 20th centuries, auto kath auto ceases to signify a determinate domain, but instead signifies a profound and deep indeterminacy inseparable from all opposition. “Determinate” comes to develop connotations that its Ancient past did not, namely connotations of irreducible relation and incompleteness. In other words, the determinacy of the Ancient world becomes penetrated by those very elements of indeterminacy that it so desperately eschewed. Because the Absolute appears to be indeterminate, the Absolute would share all of the same cognates as the indeterminate. As non-positional, the Absolute would be infinite, inconceivable, unknowable, indefinable, and ineffable. Before the post-Kantian turn to integrate the indeterminate into the body of philosophical science, it is the tradition of apophatic theology, the way of negation or via negativa, which seems to have recognized the profundity of this problem above all others. In the first deduction of the Parmenides, Plato has Parmenides deduce that if the One is, then it is unlimited (ἄπειρον) and without form.29 Following Plato’s insight that the One is indeterminate, Plotinus argues that the One emanates (απόρροια) or overflows out of itself, thereby giving forth the world of determinations. Plotinus, following Plato, recognizes the positive significance of the indeterminate: in virtue of overstepping its own limitation, the One not only makes possible but also begets all determinations. Meister Eckhart, for example, recognizes that “the divine being is equal to nothing, and in it there is neither image nor form.”30 Just as holding indeterminacy apart from all determinations shows its inseparability from all determinations, Eckhart recognizes that God (the Absolute) exceeds his own limits, namely that he is “free of all things, and therefore he is all things.”31 For Eckhart, the Absolute is ecstatic—it stands outside itself. The Neo-Platonic tradition from Pseudo-Dionysius all the way back to Plato is intimately engaged with indeterminacy as it pertains to the Absolute. As Plato’s (much disputed) Seventh Letter has it, the first principles of human knowledge cannot be fully expressed by means of language: There is no treatise (suggramma) by me on these subjects, nor will there ever be. For it is by no means statable (rhe-ton) like other topics; but from much study of the subject and from living with it, suddenly, as from a leaping fire, a light comes to be in the soul which now feeds itself.32

Emerging Recognition of the Significance of Indeterminacy  11 What is more, Plato himself posits the indeterminate as one of four principles of the universe in Philebus.33 Aristotle famously claims that Plato invoked two principles: the One and the indefinite [or indeterminate] Dyad, the latter of which is constituted by the Great and the Small.34 Plato’s significant insight on the dyadic structure of the indeterminate seems to track two significant features of the indeterminate. If we speak in terms of magnitude, it seems to evade all definite positions, and appears to be too small to be measured (which in calculus is called the infinitesimal). On the other hand, it disappears into the whole as infinitely great: it’s magnitude is too large to be measured. Following Plato, in his Elements of Theology, Proclus posits the ἄπειρον and the Πέρας as two of three principles that follow from the One. Although Aristotle distinguishes himself from Plato and all other thinkers when he says that “everyone makes all things come from contraries, but neither the ‘all things’ nor the ‘from contraries’ is right, [. . .]”35it is Plato, more than Aristotle, who seems to anticipate the positive significance of indeterminacy. In a millennium in which determinacy is privileged over indeterminacy, Anaximander’s apeiron survives on the fringes of the Western tradition in Neo-Platonic philosophy. To be sure, it is revived and brought into the body of philosophical science in post-Kantian philosophy. As we have pointed out earlier, traditionally determination has been privileged over indeterminacy as concerns the form of the Absolute. This is evident in Kant’s second principle of determination, the principle of thoroughgoing determination. Moreover, this principle also illustrates the connection between the Absolute and determinate thinking, as well as the importance of modal considerations. The view that all things are determinate is expressed in the proposition that “everything exists is thoroughly determined.36” For Kant, this is justified by appeal to the (non-formal) principle that as concerns every concept of a thing, “among all possible predicates of things, insofar as they are compared with their opposites, one must apply to it.”37 According to this principle, each concept of a thing is completely determined by comparison with the whole of possibility or “the sum total of all predicates of things in general” from which one restricts the possibilities which might apply to it. Moreover, the procedure by which any particular concept is thoroughly determined works by means of a disjunctive syllogism.38 For simplicity of communication, we can assume for the sake of clarity that the whole of possibility is constituted by two possibilities: “A” and “B.” The subject “s” would become thoroughly determined in the following way: Either A or B is predicated of s. A is not predicated of s. Thus, B is predicated of s. Thus, the possibility of the concept of each thing (as completely determined) is grounded in the predicates it acquires from the totality of possible determinations.39 S becomes determinate because the possible predicates that apply to it are limited, and determination involves positing a limit. The determination is complete because all the predicates that apply to S are attributed to it. Following Kant, the concept of the total determination of a thing appears to be inseparable from modal questions about possibility.

12  Gregory S. Moss, C.U.H.K. Long before Heidegger’s famous critique of onto-theology, Kant offered his own critique of onto-theology grounded in the principle of thoroughgoing determination [Grundsatze der durchgängingen Bestimmung]40. By positing a Transcendental Substratum under the sum of all possible predicates, one arrives at a concept of a being, namely God, which becomes the subject of the sum of all possible predicates.41 If the thoroughgoing determination of things is grounded on a transcendental substratum, reason arrives at the concept of the All of Reality or the omnitudo realitatis.42 The condition of the possibility of the complete determination of things, is thereby transformed into the Transcendental Ideal, a concept of a Thing-in-Itself (a thing independent of all possible experience), which has been variously described as an ens realissimum (the most real being), ens originarium (the original being), ens summum (the highest being), and ens entium (the Being of beings).43 Although reason does presuppose the Idea of such a being,44 and this Ideal represents a relationship between an Idea, a representation of the Absolute, with concepts,45 the Ideal, the Being of beings, cannot be given determinately: Now if we pursue this idea of ours so far as to hypostatize it, then we will be able to determine the original being through the mere concept of the highest reality as a being [. . .] in a word, we will be able to determine it in its unconditioned completeness through all predications. Meanwhile, this use of the Transcendental Ideal would already be overstepping the boundaries of its vocation and its permissibility. For on it, as the concept of all reality, reason only grounded the thoroughgoing determinacy of things in general, without demanding that this reality should be given objectively, and itself constitute a thing. The latter is a mere fiction, through which we encompass and realize the manifold of our idea in an ideal, as a particular being; for this we have no warrant, not even for directly assuming the possibility of such a hypothesis.46 Indeed, Kant’s insight is profound: the principle of all determinations cannot itself be a determination. Kant is clear that there is no object congruent to this Ideal;47 it can never be given in concreto.48 Indeed, the Idea of the Absolute, the principle in virtue of which there are determinations, must remain indeterminate. Despite their very different philosophical approaches, just as for Kant the Being of beings, i.e., the Transcendental Ideal, is not itself a being; for Heidegger, Being is not a being. Heidegger points out that the indefinable character of Being only means that Being cannot have the character of an entity.49 In a famous passage, Heidegger claims that, The Being of entities “is” not itself an entity. If we are to understand the problem of Being, our first philosophical step consists in not muthon tina diegeisthai, in not ‘telling a story’—that is to say, in not defining

Emerging Recognition of the Significance of Indeterminacy  13 entities as entities by tracing them back in their origin to some other entities, as if Being had the character of some possible entity.50 Just as Being is always only the Being of beings, for Being is not a separate thing apart from beings, we are only ever on the way to Being, for it is just this absence of being as a thing that sets us on a course without end. Philosophy as Free Science Having formulated the problem of conceiving the indeterminate, we might re-formulate the opposition of indeterminacy with determinacy: while the former appears to be by itself, the latter is in relation. What is by itself is independent and autonomous. Accordingly, an important locus of the investigation into indeterminacy involves the concepts of independence and autonomy. We should remember that Ancient Greek philosophy identified determinate existence with the property of independence expressed in the phrase auto kath auto, itself by itself. As philosophy comes into the 19th and 20th centuries, auto kath auto ceases to signify a determinate domain, but instead signifies a profound and deep indeterminacy inseparable from all opposition. “Determinate” comes to develop connotations that its Ancient past did not, namely connotations of irreducible relation and incompleteness. In other words, the determinacy of the Ancient world becomes penetrated by those very elements of indeterminacy which it so desperately eschewed. Although Kant’s philosophy appears to privilege the determinate as the form of cognition, cognition would not be possible without indeterminate principles, such as the indefinite Mannigfaltigkeit and the Ding An Sich or Thing-In-Itself. In regard to the former, cognition requires an indefinite intuitive content that becomes cognizable in virtue of being determined by the categories.51 Since only possible objects of experience can be known, and the indefinite manifold is not itself an object of experience, the indefinite manifold cannot be known. Hegel notices the paradox that although the indefinite manifold is unknowable, it must be known as condition for the possibility of experience. Likewise, the Thing-In-Itself is indeterminate,52 for it transcends all possible experience, and can neither garner form from categories nor content from intuition. Yet, the Noumenon is a posit that is necessary for maintaining Kant’s distinction between phenomenon and Noumenon. Hegel and Husserl recognize that in order for knowledge to be possible, the indefinite must be introduced into the body of knowledge per se. In other words, the indeterminate cannot be known as a condition of knowledge without being constitutive of knowledge in some way. This turnabout in the conception indeterminacy marks a turning point in the tradition concerning the very conception of philosophy. On the one hand, if rational knowing is constituted in a fundamentally determinate way, then rational knowing would be insufficient for uncovering the truth of

14  Gregory S. Moss, C.U.H.K. indeterminacy and its related determinations: the Absolute and autonomy. We might designate this non-discursive path to uncovering the indeterminate as existential, and it connotes one of the forks in the Western tradition. On the other hand, rather than give up on rational knowledge of the indeterminate, philosophy might instead re-think the very structure of knowing. Following Aristotle, both Hegel and Husserl contend that philosophy ought to be an autonomous science of wisdom. Unlike Aristotle, autonomous thinking can only become autonomous if philosophical science becomes non-positional, namely if it is not beholden to any privileged position from which it might begin. Philosophy can recover the positive significance of indeterminacy by construing the indeterminate as transcending philosophical knowing, where the latter is a constraint or negation of the former. On the other hand, philosophical knowing can recover the positive significance of indeterminacy by introducing indeterminacy into the very body of rational knowledge from which it has traditionally been excluded. Simply put, indeterminacy can overcome its overwhelmingly negative significance by becoming existential or free science (in the sense of Wissenschaft). Robert Scott, in Indeterminacy As Key to a Phenomenological Reinterpretation of Aristotle’s Intellectual Virtues, brings Aristotle into dialogue with Husserlian phenomenology. There, he challenges the Aristotelian prioritization of determinacy according to which the highest intellectual virtue for Aristotle (philosophic wisdom) consists of immaterial, determinate knowledge of invariable truths. Scott argues that a phenomenological retrieval of the Aristotelian intellectual virtues can both renew the latter and show how the positive significance of indeterminacy can be recovered as integral to philosophic wisdom. Accordingly, the Kehre in the conception of indeterminacy correlates with a transformation in the conception of what constitutes free science. Free science must proceed from an indeterminate position from which determinate concepts and structures can be further developed. In the language of Hegel53 and Husserl, the demand is that philosophy ought to be without presupposition (Voraussetzung), of which a more literal translation might be “without an advanced (voraus) position (Setzung).” Of course, the way each attempts to achieve this is vastly different: Hegel’s Wissenschaft develops a purely conceptual (or “speculative”) approach, whereas Husserl’s science is that of a descriptive Phenomenology. The so-called Myth of the Given might here be construed as the myth of absolute positionality. The positivity of indeterminacy is illuminated nicely in Hegel’s dialectic, and in Husserl’s concepts of ἐποχή and horizonality. To put it simply—indeterminacy is indissolubly linked with questions of beginnings—whether these be philosophical or otherwise. To begin with Hegel, we discover a contradiction in thinking about determinacy in absolute terms. When Hegel describes determinateness as “negation posited as affirmative,” he indicates that the whole structure of something/other is formulated in terms of difference and negation. Seen in

Emerging Recognition of the Significance of Indeterminacy  15 this light, for one to limit the form of logical cognition or every possible truth to some determinate form is to privilege the reality of “negation” as an absolute or non-relative principle. A contradiction arises: being determinate or in-a-position cannot be non-relative or absolute without abandoning all particular positions. Thus, the problem of indeterminacy arises through the treatment of determinacy as an absolute or non-relative standpoint. In other words, if one cannot think indeterminacy without raising it to the position of the Absolute, one cannot privilege determinacy as an Absolute measure without contradiction. Hegel’s Wissenschaft der Logik begins with Being [Sein]: “Being is the indeterminate immediate.54” It is “pure indeterminateness.”55 As indeterminate, it cannot be posited vis-à-vis nothing, and cannot be distinguished from it: “each of them [Being and Nothing] is in the same way indeterminate.56 Just as nothing is seen in pure darkness, so can nothing be seen in pure light. Hegel is careful not to fall into an onto-theological way of thinking, for he carefully distinguishes Sein from Dasein—the latter is determinate being, not mere Being. Hegel is clear that “Determinate Being is the first category to contain the real difference of being and nothing, namely, something and other.”57 Since free science begins with indeterminate Sein, then determinate being, Dasein, must develop out of the indeterminate. Given that determinate being and knowing must arise out of indeterminate being and knowing, Hegel points out that philosophers have “placed the essence of philosophy in the answering of the question: how does the infinite go forth from itself and make itself finite?”58 Although Hegel discusses this question in respect to the infinite, the fact that the indeterminate has the infinite as one of its primary cognates is sufficient for us to introduce it here. Hegel comments that whether there can be a philosophy at all is thought to depend upon this question.59 Indeed, it is only because of the recognition that philosophy cannot begin with determinacy that the problem of how the determinate arises in the first place becomes of central issue. Put in metaphysical terms, this is one of the most familiar questions: why is there something instead of nothing? Ex nihilo nihil fit stands as a testament to the hegemony of determinacy. Given the importance of this question for a philosophical treatment of indeterminacy, Ricki Bliss and Filippo Casati address this question in Cosmological Questions by confronting the multi-vocal senses of nothing and its ambiguity in a comparative context. In addition, Robert Neville, in Indeterminacy in Chinese Thought: Spontaneity and the Dao links themes of spontaneous beginnings and the question concerning the possibility of a world by illustrating how the Dao spontaneously creates something out of nothing. For Aristotle, free thinking is constituted by νοῦς, in which thinking never oversteps its own limit, but rather remains at the place of its own limit: ἐντελέχεια. Following Hegel, if free science must integrate the indeterminate, and thereby become non-positional, free thinking can no longer remain at its limit as with ἐντελέχεια. Rather, it must do the very opposite: overstep

16  Gregory S. Moss, C.U.H.K. its own limit. For the non-positional is that which exceeds its own limit: in virtue of standing opposed to positionality it ceases to stand opposed to determinate being and thereby becomes what it is not. Aufhebung is constitutive of free thought, for it is the activity by which thought exceeds its own limit. Sein, the indeterminate, exceeds its non-positionality in order to become determinate or positional in Dasein. Indeed, non-positionality does not have a position in the system of knowledge until it is brought in relation into Dasein.60 In other words, the indeterminate can only acquire the position as the first category once it has exceeded itself and given rise to a second with which it can stand in contrast to as its beginning. Indeed, the self-exceeding power of the indeterminate has immensely positive significance for philosophical method. For Aristotle, the Noumena or first principles are thought by means of Nous, each of which is conceived as determinate. Kant, having brilliantly recognized that what is properly autonomous cannot be rendered determinate by cognition, relegated the Noumena to the unknown. Having banished the Noumenon to the indeterminate beyond, Kant banishes Nous61 as well from the purview of human knowing. By introducing the indeterminate into the content of philosophical science, Hegel means to develop a science of Noumena. But Hegel recognizes that this science of Things in Themselves requires a radical transformation of both Nous and Noumena. To think the indeterminate Noumenon, Nous must be constituted by indeterminacy—what exceeds its own limit—and become Aufhebung. Because the science of logic begins with the indeterminate, for Hegel indeterminacy has the positive significance of constituting the beginning of philosophy. Perhaps more than any other, Hegel is profoundly sensitive to the dialectical features of indeterminacy. Since the whole science is the activity by which the indeterminate exceeds itself into various determinate forms, the whole of the science can be characterized as various determinations of indeterminacy. Again, since every category of the logic is a definition of the Absolute, or what is by itself, and the latter is indeterminate, the indeterminate is not maintained as one category alongside others, but is integral to the content of the other categories that are developed. Indeed, that the non-positionality of indeterminacy renders it indistinguishable from the various categorial positions in the logic can be gleaned from a brief gander at the text. The One is “indeterminate but not, however, like being, its indetemrinateness is the determinateness which is a relation to its own self, an absolute determinateness.”62 Essence is “indeterminate simple unity from which what is determinate has been eliminated in an external manner.” Or more starkly: “Absolute Essence [. . .] has no determinate being.”63 In regard to Mechanism, Hegel points out that “the object is therefore in the first instance indeterminate insofar as there is no determinate opposition in it.”64 As is evident, in Hegel indeterminacy acquires the enormously positive significance that the various determinations of the Absolute, such as One, Essence, and Mechanism, are all forms of indeterminacy.

Emerging Recognition of the Significance of Indeterminacy  17 Integral to dialectic’s self-exceeding is the correlate concept of absolute determinacy. In exceeding itself, the non-positional becomes positional and thereby excludes no determination. Every determination, including determinacy and indeterminacy, have been cancelled and preserved. Since this non-positional character is not devoid of determination, but includes them all, Hegel calls it absolute determinacy in which the negation that constitutes determinacy has been negated, and is expressed in Hegel’s famous expression of the negation of negation. Take two examples: “Being for Self” is an infinite determinateness, because it has absorbed all distinction.65 “The Concept” is an absolute determinateness.66 For Hegel, the negation of negation is a formula of the logic of freedom. Although initially absolute and indeterminate, freedom gives itself a relative and determinate form. Finally, freedom negates its own form of determination, thereby re-establishing its absolute character, though now in determinate form, as the negation of negation. Just as Aristotle conceived of free thought as the contemplation of thought thinking itself, νοησις νοησεως νοησις, so Hegel and Husserl conceive of free science as a science of self-thinking thought. Whereas for Aristotle selfthinking thought is radically finite and eschews indeterminacy, for Hegel and Husserl it contains indeterminacy, and is infinite. On the one hand, Hegel conceives of free science in the Wissenschaft der Logik as purely logical in character for which the Phenomenologie des Geistes is only a mere propaedeutic. Rather than follow Hegel in his complete abandonment of the opposition of consciousness as the foundation of epistemology, Husserl re-thinks the structure of self-consciousness in such a way that incorporates indeterminacy into the method and the content of the philosophical science of self-consciousness. The freedom of phenomenology can first be gleaned from the ἐποχή. Theorizing from presuppositions (Voraussetzungen) takes its start from the natural attitude. Among other things, in the natural attitude I relate to independent things in an immediate way.67 The immediate signifies absorption in the object of consciousness, such that the structure of consciousness itself remains hidden. In this attitude it is obvious that the world itself and all its inhabitants, e.g., values, customs, bodies, etc., is immediately accessible to me. In this attitude I relate to myself just in this same way: as a being in the world. When I begin to think philosophically, I do not necessarily abandon this attitude by ceasing to live unreflectively. Rather, from within the natural attitude my attention becomes directed to a different set of concerns and objects. This kind of thinking begins from prejudice: we posit particular beings (vorgegebene Seinende) as given evidence from which our reflections proceed.68 Thinking becomes free from the natural attitude in epoché. The Pyrrhonian skeptics employed this term to signify the suspension of judgment by which one achieved peace of mind. Epoché puts out of action “any thesis which belongs to the essence of the natural standpoint, we place in brackets whatever it includes respecting the nature of Being.”69 Epoché sets it out of action and disconnects from it.70 By setting it out of action, the

18  Gregory S. Moss, C.U.H.K. phenomenologist sets all independent positions that are given in advance out of action. Accordingly, philosophical knowledge becomes free from all presuppositions and prejudices (Voraussetzungslosigkeit und Vorteilslosigkeit).71 As he puts it in his Philosophy as Rigorous Science (Philosophie als Strenge Wissenschaft) philosophy becomes a proper science only when we no longer take anything for granted in advance.72 Epoché makes free science possible by making no use of the natural attitude and thereby frees itself from presuppositions. Although this act of bracketing sets the indeterminacy of the natural attitude out of play, in what follows it will become more evident that it is just as much a work of indeterminacy. What is more, the re-evaluation of the role of the natural attitude in epoché is a “concern of our full freedom”73 and this freedom cannot be won without the power of indeterminacy. Because bracketing suspends judgment regarding all positions given-in-advance, bracketing does not take a position in regard to the natural attitude. Thus, epoché is an act of indifference or Gleichgültigkeit (to use a term from Schelling) toward the natural attitude. Indifference neither affirms nor denies the object to which it is indifferent: rather it does not concern it whatsoever. Accordingly epoché is not positionally related to the assumption(s) that it brackets, for it neither takes the position of affirmation nor negation. Because I relate to thinking as positing in a non-positional way, and the non-positional is what constitutes indeterminacy, the act of bracketing institutes a fundamentally indeterminate relation to the naiveté of the natural attitude in its everyday and reflective stances. Naturally, the epoché is a position, just like Pyrrhonian skepticism, but it is a position that fails to take a particular position vis-à-vis the natural attitude. In sum, indeterminacy has positive significance for phenomenology, for epoché frees the phenomenologist from presuppositions in virtue of the indeterminacy inherent within it. The freedom of phenomenology is not just a negative freedom, but a positive freedom that wins an autonomous domain of inquiry. By bracketing the immediate relation of consciousness to the independent thing, the field of consciousness of is opened to philosophical investigation: “thus we fix our eyes on the sphere of consciousness and study what it is that we find immanent in it.”74 Rather than investigate things independently of consciousness, phenomenology inquires into itself, and is a science of self-consciousness, or Selbstbesinnung.75 Consciousness has its own being that is absolutely unique and remains unaffected by epoché.76 By freeing philosophical science from presuppositions, philosophy wins an autonomous field of inquiry: The spirit and, in fact, only the spirit is a being in itself and for itself; it is autonomous and is capable of being handled in a genuinely rational, genuinely and thoroughly scientific way only in this autonomy.77 The indeterminacy afforded by the epoché not only frees the philosopher from presuppositions, but also has positive significance methodologically

Emerging Recognition of the Significance of Indeterminacy  19 insofar as it opens up and maintains the domain of self-consciousness to phenomenological inquiry. Although Kant had already called into question the naiveté of the natural attitude in philosophy, his critique presupposed categories that were not themselves subject to transcendental investigation. In order to proceed without presupposition formal logic as well as the indeterminate manifold must become an object of inquiry and cannot thereby be given in advance. Accordingly, through bracketing, the consciousness of objects could only be illuminated by means of description rather than deduction from given formal principles. Rather than generate all determinacy out of indeterminacy as Hegel attempted, phenomenology suspends all given presuppositions in order to attend to what is there for consciousness. Although the supposition of the independent thing is suspended, objects do appear in relation to consciousness. As is well known, Husserl calls the consciousness of structure intentionality a transcendental structure by which objects are given to and constituted by consciousness. Consciousness is conscious of an object, and refers to that object by means of the two-fold structure of noesis and noema. The noema is the sense or meaning in virtue of which an object is given, e.g., the reference to Venus as “the morning star” or what might be termed the content of the intention, and the noesis is the way that the sense of an object is intended, such as by judging or imagining. Since what is left over after the bracketing is the consciousness of objects, whatever phenomena are encountered are constituted by consciousness. Accordingly, no limiting principle can be discovered from within the phenomenological approach by which the independence of transcendental consciousness could in principle be limited. Indeed, Husserl often describes the field of phenomenological description as absolute and autonomous. Husserl posits the transcendental Ego as the source of the intentional awareness of objects, to which every “consciousness of” points back, and in this sense all experience and the field of knowing is relative to self-consciousness, which is absolute and autonomous. The objects themselves are constituted by consciousness: they are not beyond it. For example, my perception of this table involves the noesis of perception, and the noema “table,” by which this table is given. I experience this table as one self-same unity that recurs throughout the consistent changes in the adumbrations of my experience. Every image of the table in my perceptual experience is an image of the self-same table. In suspending all prejudicial knowing, and opening up the eidetic description of its noema, phenomenology is attending to the things themselves insofar as they have meaning, as there is no other meaning to which one can in principle attend. Husserl’s famous slogan “back to the things themselves!” is reflected in employment of the term noema. The noema in Greek thought is the thing which one comprehends in Nous, and exactly what Kant barred as a possible object of cognition. Although Husserl rethinks the meaning and significance of the thing itself, Husserl recognizes that a return to the

20  Gregory S. Moss, C.U.H.K. things themselves cannot be completed without a return to the indeterminacy that accompanies all knowledge of noema. Indeed, it is in the concept of the horizon that we discover an indeterminate condition for the givenness of any object. Whatever place we inhabit is enclosed by a horizon that has a position in experience while simultaneously transcending all places. Indeed, the importance of the phenomenology of place is bound up with the question of indeterminacy, and must come to terms with the issue of the horizon. With the return of the thing-in-itself within awareness, and the introduction of the indeterminate manifold into the object of intentionality, the indeterminate cannot help but appear within phenomenological consciousness in the form of the horizon. Horizon stems from the ancient Greek ὁρίζων, which signifies to “mark out a boundary.” Following our conceptual analysis, horizonality would signify some kind of determinacy for the determinate as a limiting concept. Ordinarily, horizon connotes the line where the sky meets the ocean. Nothing appears beyond the horizon. Accordingly, horizons limit and thereby determine what can and cannot appear in any particular experience. For this reason, the horizon is that which enables phenomena to appear. Although the horizon appears to be a principle of determinacy, it is itself relative and indeterminate. The horizon does not have an absolute position. Where the horizon appears depends upon the situation of the subject of experience. Wherever I happen to be, the horizon cannot be. It is always at the fringe of the field of experience (whether understood psychologically or phenomenologically). The relativity of the horizon indicates its indeterminacy: as I approach it, it recedes. Accordingly, it is a limit that cannot be surpassed; it is a line that cannot be crossed. As I approach the position of the horizon, its position changes: it’s absolute position is indeterminate. Whenever I arrive at the place of the horizon, the horizon has shifted. In his A Note on the Phenomenology of Horizons, Crowell investigates the various ways that Husserl, Heidegger, and Merleau-Ponty characterize the concept of the indeterminate horizon. Because the horizon is the limiting principle that allows phenomena to appear, every noema has horizons: both internal and external. As I perceive the cover of A Room with a View, I perceive the book, but not in its completeness. The book has a horizon in which it appears, such as the back side of the book which I do not see, but makes possible the appearance of the cover of the book as part of the same book. Likewise, the book is perceived to appear within an outer horizon, namely on a table in a small room.78 Further, the book appears in horizons that may not be given in the field of vision, such as the home, the neighborhood, the city, nation, and world. Since noema appear to consciousness within horizons of meaning, the intention of some noema is also always an intention of some horizon of meaning, what Saulius Geniusas calls “co-intentionality.”79 Though indeterminate, horizons seem to function both as an indeterminate object of intention as well as conditions for the possibility of intention.

Emerging Recognition of the Significance of Indeterminacy  21 Given that sense and horizon are bound together, the origin of sense cannot be accounted for without investigating the origin of the horizon. Accordingly, the concept of horizon is bound up with Husserl’s concept of genetic phenomenology. Whereas static phenomenology investigates consciousness in the present, genetic phenomenology investigates the origin of that present temporally, historically, and structurally. Indeed, the motivation for genetic phenomenology may be motivated by the appearance of an uncanny horizon at the limit of all static analysis80 and in this way carries positive significance. Janet Donohoe’s article, Genetic Phenomenology and the Indeterminacy of Racism further elaborates on Husserl’s genetic phenomenology in regard to the themes of home and alien world. Donohoe’s chapter calls for significant ethical reflection on the sedimented, historical sources of ideas, attitudes, and institutions (such as racism). Although the horizon is an indeterminate limit, I can determine what was the horizon for an object, and constitute it as a determinate object of intentionality. Nonetheless, since the horizon always again pivots to the fringe, I can never exhaustively determine the horizon. Although the horizon appears determinable, it cannot be exhaustively determined. Unlike Hegel’s concept of absolute determinacy, in which philosophical thinking of the absolute can be completed, phenomenology remains an infinite task. Given that the horizon cannot be exhaustively determined, yet is always co-intended in experience, experience is saturated with an inexhaustible indeterminacy. This becomes most evident when we reflect on the world-horizon or the horizon of all horizons. As I approach the limit of consciousness, I approach the horizon of all phenomenological horizons. Since it always recedes upon approach, the position of the limit of phenomenological consciousness is irreducibly indeterminate. Because I cannot set the world-horizon into a determinate relation with other noema, for it encompasses them all, what makes any phenomenological field accessible and determinable is an absolutely indeterminate limit—an ὁρίζων. Having bracketed the natural attitude, self-consciousness opens as an infinite field of description, whose absolute limit is an indeterminate, unsurpassable horizon. The horizon offers a sense of indeterminacy that is fundamentally positive in significance: it makes possible the manifestation of phenomena. Beyond Philosophical Science: Existential Indeterminacy Rather than incorporate indeterminacy into free science, an alternative path invokes the non-discursive character of intentionality. In contemporary philosophical discourse, in good Aristotelian fashion, Markus Gabriel attempts to avoid this result by re-asserting the priority of the determinate over the indeterminate with his provocative book Sinn und Existenz, in which he argues that the world does not exist. To put the problem in our terminology,

22  Gregory S. Moss, C.U.H.K. since the world is Absolute, it must be absolutely indeterminate, and must be non-positional. Gabriel defines existence in terms of determinacy: to exist means to have some (minimal) position, or in Gabriel’s terms to be located (vorkommen) in some field of sense (Sinnfeld). Since the world has no position, and to exist is to have a position, the world does not exist, it is a false nothingness.81 Note that Gabriel is not denying the existence of all indeterminacy, just absolute indeterminacy. Gabriel’s contradiction is another version of our problem of indeterminacy: the position of the world is its non-positionality. Rather than regulate the absolutely non-positional to non-existence, one could follow Hegel and Graham Priest by rejecting the principle of explosion and argue that there can be true contradictions. Since the non-positionality of the indeterminate certainly leads to a contradiction, namely both F and not-F, in order to affirm the truth of the indeterminate one might argue that contradictions can be true. Although the form of Graham Priest’s logic (paraconsistent logic) by which he defends his theory of truth (Dialetheism), is formal (while Hegel’s dialectic is not), both Hegel and Priest can give an account of the possibility of a science of the indeterminate. In Heidegger and Dōgen on the Ineffable, Priest and Filippo Casati give a Dialetheist reading of Dōgen and Heidegger, and show that Nishitani’s conception of absolute nothingness undergirds the contradiction in both Dōgen and Heidegger’s conception of the ineffable. As we think through these possibilities, let us follow what Dickman suggests in his Hermeneutic Priority and Phenomenological Indeterminacy of Questioning, namely that to practice genuine questioning, which is dialogical, opens us to new possibilities and calls us to constantly rethink what is possible. But there is yet another tendency: the tradition of the Kyoto School, Existentialism, and Post-Structuralism (to take a few visible examples), which acknowledge the indeterminate (and in some cases, assert that the indeterminate exists), but either problematize any complete scientific grasp thereof (such as Deconstruction) or problematize any discursive account as violating some privileged logical form. If the form of knowledge is fundamentally positional, and precludes the possibility of contradiction, in order to access the indeterminate condition for determinate knowledge claims, philosophy must transcend the limits of conceptual knowing, and come to terms with the insurmountable gap between philosophical knowing and the indeterminate. In this sense, philosophy becomes a crisis in the sense of κρίσις—it is a crisis constituted by the separation between indeterminacy and philosophical knowing. As Stephen Barnett astutely notes, “Hegel defines the modernity that our postmodern era seeks to escape.”82 Whether it be Lyotard, who lambasts Hegel as the “philosopher of homogeneity,”83 or Levinas in Totality and Infinity, who attempts to rescue the Other from Hegel’s dialectic of totality, a dominant strain in Continental philosophy can be characterized as resistance to Hegel’s transformation of indeterminacy into absolute determinacy.

Emerging Recognition of the Significance of Indeterminacy  23 Derrida recognizes that Hegel is not just the philosopher of homogeneity, but also “the thinker of irreducible difference”84 According to Derrida: If there were a definition of différance, it would be precisely the limit, the interruption, the destruction of the Hegelian dialectical synthesis wherever it operates.85 Determinate being, as we have noted, has been conceived in modern philosophy as characterized by binary oppositions. Like Hegel, Deconstruction undermines such binary oppositions. Unlike Hegel, Derrida eschews logocentrism. Deconstructive thinking unsettles any absolute position. By doing so, it certainly determines the indeterminate, and thereby differentiates it. But this differentiation is not a result in a binary opposition, but indefinitely defers the indeterminate to another difference. For every difference that is posited, the indeterminate is further differentiated and deferred. Although Hegel also aims to transcend binary thinking, in deconstruction one never arrives at the absolute determination. This incompleteness constitutive of deconstructive thinking is further illustrated in the concept of the trace, where the relation with the other is marked. Every marking of the indeterminate introduces a trace—a relation to an other which remains beyond whatever is said. Rather, unlike Hegel’s absolute determination, deconstruction is a thinking without end. In difference and deferring, différance signifies an irreducible semiotic incompleteness. Deconstruction is a thinking that does not settle in any final position. To the contrary, it unsettles every position or is the thinking that unsettles any and all absolute positionality.86 Thinking— conceived as the unsettling motion that displaces absolute positionality, certainly does not begin with Derrida—but is pre-figured in the entire history of Continental philosophy. Indeterminacy, as non-positional, becomes determinate in fixing it in opposition to determinate being. While philosophical knowing constrains what is known in a matrix of determinate self-identical being, the indeterminate exceeds its limit. What exceeds its limit becomes what it is not. To polemicize the ancient terminology, we might state that determinate being is; indeterminate becomes. In the thought of Nietzsche, we find a philosopher dedicated to reviving indeterminate becoming from the clutches of “being” or being in the form of conceptual determination. Nietzsche identifies the attempt to constrain the indeterminate by rational form conceptual idolatry: All that philosophers have handled for millennia has been conceptual mummies; nothing actual has escaped their hands alive. They kill, they stuff, when they worship, these conceptual idolaters.87 By fixing the indeterminate flux, the philosopher makes a mummy of becoming. Becoming is not the image of being; rather, being is the death of becoming. What merely is, but does not become, is a lie. Having arrived at

24  Gregory S. Moss, C.U.H.K. the awareness of this idolatry, Nietzsche tells the story about how the “real world” (the world of determinate Form) became a lie.88 To put it even more polemically: Once upon a time, in some out of the way corner of that universe which is dispersed into numberless twinkling solar systems, there was a star upon which clever beasts invented knowing.89 For Nietzsche, conceptual idolatry obscures the truth of becoming and its inherent indeterminacy. Indeed, the world of becoming, which is radically indeterminate and individual, is rendered conceptual by identifying unequal things: “every concept arises from the equation of unequal things.”90 By rendering what is indeterminate determinate, we equate what is unequal. Thus, we can only conceive the indeterminate by identifying what is not identical. Indeterminacy is not determinate, yet we render it determinate by conceiving it. If we only attend to the form of determination—the grammar of knowing, if you will,91 we often fail to see the indeterminate condition and contradiction lying at the ground of determinate knowledge. By uncovering the contradiction at the base of conceptual determination, the philosopher can free herself of conceptual idolatry. By freeing becoming from being, Nietzsche means to reclaim a positive significance for indeterminacy. For Nietzsche, of course, the positive significance of becoming is not just present in his ontology and epistemology, but is also evident in his concept of the flourishing life. Desire does not realize itself; desires can go unfulfilled. Because obstacles stand in the way of its fulfillment, affirmation of desire must also entail an affirmation of the overcoming of the resistance imposed by the obstacle. Accordingly, the happiness that comes from fulfilling desire involves the feeling of overcoming resistance.92 Indeed, a life that affirms desire (rather than eliminating or reducing it) requires constantly overcoming new obstacles. Every obstacle overcome is a way of exceeding one’s own limitations. Rather than longing to remain the same, a flourishing life is always challenging itself by bringing life’s greatest weapons against oneself.93 As Nietzsche says, rather than stay the same, one only remains fruitful by being “rich in contradictions.” One remains young as long as one does not long for peace. Nietzsche’s concept of creativity is certainly bound up with becoming. Although the human being is a “piece of fate,” and denies free will, in the “instincts” of the Übermensch, it is nature itself that creatively becomes by constantly exceeding its own limitations. Long before Nietzsche’s revival of becoming, Chinese and Buddhist philosophers both acknowledged and grappled with indeterminacy in their ontologies and systems of ethics. Qingjie James Wang, in Body, Intimateness and the Ground of Confucian Ethics, elaborates on the formative character of Confucian ethics that arise from the Confucian embrace of the body. Wang further develops the conceptual connections between the indeterminacy of the body and the Confucian understanding of intimacy and graded

Emerging Recognition of the Significance of Indeterminacy  25 love. As regards indeterminacy and the body, Crowell approaches the body from a phenomenological point of view, in his discussion of motor intentionality in Mealeau-Ponty. Crowell connects the existential core of signification, situational motoricity, with Ponty’s concept of the fundamental ambiguity of being. Buddhist philosophers in the Mahayana tradition, such as Nagarjuna in the Fundamental Wisdom of the Middle Way, radicalized the concept of indeterminate becoming. In addition to the Buddhist principle of dependent origination, “this arises, that arises,” and the view that all things are related in a cycle of becoming, this school advocated the position that all phenomena, including enlightenment itself, are empty of independent and determinate being. Keiji Nishitani, a representative of the Kyoto School, practitioner of Zen, Nietzsche scholar, and student of Martin Heidegger, further develops the concept of sunyata (emptiness) into a philosophical principle. One fundamental insight of Zen, which we have discussed at some length in our analysis of the history of Western philosophy, is the principle that autonomy is indeterminate. What is absolutely autonomous or non-positional cannot be distinguished from what is positional without the non-positional ceasing to be non-positional. Positional or determinate being is empty insofar as it fails to have an independent, determinate position vis-a-vis indeterminacy. Accordingly, the non-positional permeates every position: it is absolute, and is what Nishitani calls circuminsessional.94 Insofar as all positions are united in the non-positional, all things have Samadhi-being,95 for all things are concentrated into one. “All dharma gates are open” expresses the accessibility of the absolute from each element of the whole. Because all determinate beings are concentrated into the indeterminate, they are concentrated into nothing at all, for the indeterminate is nowhere and no-thing. Thus, insofar as they are concentrated into one, they fall apart and are dispersed into an indefinite plurality. For these reasons, although all things are related, they are related in virtue of the indeterminate. Thus, all things are constituted by a fundamentally indeterminate relation. Indeed, the positive significance of indeterminacy is visible in the omnipresence of indeterminacy in all relations. Given that indeterminacy is the negation of determinacy, it is a relative negation, for it is not determinacy. But it is also absolute, for despite its negation of determinacy, it is constituted by that which it negates, namely determinacy. Accordingly, indeterminacy is absolutely nothing: it is self-negation or the negation of negation. In other words, indeterminacy is empty even of indeterminate being: what lacks a position (emptiness) is empty (does not itself have a position). Insofar as the indeterminate falls into its non-being, its presence becomes indistinguishable from its very absence. The meaning of impermanence is constituted by the disappearance of the thing into its non-being, in which it does not leave a single trace behind.96 In the impermanence of indeterminacy, “the form of things scatters and falls apart.”97

26  Gregory S. Moss, C.U.H.K. Since all things are given in an indeterminate relation, the positional exclusion of one thing by another is an illusion. From the standpoint of reason, subjectivity presents itself as excluding non-subjectivity from its identity. Insofar as something is posited as something determinate, it is posited as distinct from and excluding its non-being. Since no-thing at all underlies this determinate positing, it is an appearance of nothing at all. This appearance is a phenomenon that is illusory, for it is not really separate, although it may be presented as such to sensation and reason. For Nishitani, insofar as personal being is posited as a separate and determinate self that is distinct from indeterminate relations, it is an illusion.98 The illusory character of things is constituted by the falsehood of their self-presentation as separate, independent beings, which can be first apprehended in the practice of Zazen. This is further illuminated by Nishitani’s analysis of the person as a persona. The persona is a “face” one puts on to play a role. The person is a persona. The personal, like all things, is not a persona of anything; the person is a persona of absolute nothingness. All thinking, feeling, and action are illusory appearances with nothing behind them.99 Accordingly, the self, like all things, is a play of shadows across a stage of nothingness.100 Although the illusory character of determinacy can be experienced in silence, it can only be expressed in the form of the paradox,101 such as “the indeterminate is determinate.” Rather than know the indeterminate impermanence by means of reason, we must transcend reason altogether: We are able to touch that reality only at a point cut off from the judgment and contemplation proper to reason.102 Indeed, given that the absolute is absolutely non-positional, no positional relation to the absolute will suffice to reveal what it is. Accordingly: We straighten ourselves out by turning to what does not respond to our turning, orienting ourselves to what negates our every orientation.103 Since sensation and reason are constituted by positional awareness, seeing the Samadhi being requires the practice of Samadhi, or a non-conceptual attentiveness to what is in Zazen. By letting go of the separate and determinate self, the self becomes indeterminate: a non-positional self, and thereby truly autonomous self. In the non-positional self, everything other than the self comes to constitute the position of the self: Hills and rivers, the earth, plants and trees, tiles and stones, all of these are the self’s own original part.104 Nishitani characterizes the realization of not-self (anatta) as the practice whereby the relative negation that constitutes the determinate self is negated

Emerging Recognition of the Significance of Indeterminacy  27 and thereby becomes absolute nothingness. Although both Hegel and Nishitani characterize the absolutely indeterminate as the negation of negation, the former reads this formula as an expression of the absolute determination of self-thinking subjectivity, while the latter reads it as the complete abdication of determinate subjectivity, as absolutely indeterminate not-self. Whereas Hegel defines the concept as the universal that is an element of itself, for Nishitani this very same definition is applied to the structure of non-conceptual indeterminate relations which constitute all things. In The Nietzschean Bodhisattva—Passionately Navigating Indeterminacy, George Wrisley creatively brings Nietzsche into conversation with the Mahayanan conception of the Bodhisattva ideal and argues that the latter may be compared to and perhaps enriched by a Nietzschean conception of the higher type. The revival of becoming in Nietzsche and Nishitani signals the rise of indeterminacy. Unlike Nishitani’s positioning of autonomy in indeterminate relation, in his early Being and Time Heidegger re-conceives the indeterminacy of autonomy in terms of indeterminate self-relation. In Husserl’s Crisis of the European Sciences, Husserl laments and identifies the crisis of Europe as the demise of Enlightenment rationalism.105 Indeed, it is Husserl’s own student, Heidegger, who most clearly embodies the spirit of the age of crisis in the early 20th century. Rather than draw the indeterminate into the body of science, Heidegger’s rejection of onto-theology signifies a departure from a scientific or philosophically scientific treatment of the indeterminate altogether. Among other insights, Heidegger points out that the indeterminate character of Being is itself a positive phenomenon: However much this understanding of Being may fluctuate and grow dim, and border on mere acquaintance with a word, its very indefiniteness is itself a positive phenomenon which needs to be clarified.106 Heidegger’s term for “indefiniteness” is Unbestimmtheit, which can be better translated as “indeterminacy,” what is not bestimmt or determined. Indeed, Heidegger’s rejection of onto-theology is a call to confront the evasiveness of indeterminacy as a positive phenomenon. Just as the hammer reveals its function in ceasing to work, so reason is revealed as that which disappears in the presence of the indeterminate. The placeless-ness of the indeterminate is disclosed or revealed to the inquirer only in its absence from the presence of every discursive place. To whom is the placeless-ness hidden? Every attempt to position the indeterminate within some co-ordinate system of philosophy or science induces the indeterminate to continually withdraw. Wherever reason dominates the indeterminate withdraws. To whom is this placeless-ness revealed? It is revealed to the being for whom its own Being is a question, what Heidegger calls Dasein. Unlike Hegel’s use of the word, Dasein does not have the logical significance of merely determinate being. Rather, the “Da” is the place of revelation, the

28  Gregory S. Moss, C.U.H.K. place of Lichtung.107 By looking for indeterminacy discursively, one fails to discover it: one errs in the sense of wandering on an Abweg—one goes off the path. Heidegger’s Holzweg is a kind of Abweg: by taking the wrong turn on a forest path, one comes to the place of revelation (the clearing) where the indeterminate is revealed as that which evades the discursive place. Ironically, it is in Dasein’s erring where the non-discursive presence of the indeterminate appears. The very same Dasein to whom the indeterminate remains hidden, namely the philosopher, is also the one to whom it is revealed. By seeing that reason fails to capture the indeterminate, the presence of the indeterminate is revealed to reason as that which is hidden from reason. That which remains hidden from reason is revealed as that which can only be encountered non-discursively, or existentially. Dasein is that being from whom its own being is a question. In order to make its own being visible and present to itself, Dasein must comprehend itself in its totality. To make itself visible in its totality requires understanding the limit of its existence, where it comes to an end. Dasein comes to an end in death. Thus, to understand itself, and fulfill its potential, Dasein must realize its capacity for death. Dasein is “sein zum Tode.” Of course, death cannot be realized as a potential to be, for were death realized as a potential for Dasein to be, then Dasein would no longer be. Thus, death cannot be fully present as a realized actuality. Dasein does not realize its capacity for death by dying. As Heidegger points out, death “gives Dasein nothing to be ‘actualized.’ ” Rather, it is “the impossibility of every way of comporting oneself toward anything.” It is the measureless possibility of the very “impossibility of existence.”108 Death is an unsurpassable (temporal and existential) horizon of horizons. Since death is always one’s own death, death individualizes Dasein. Accordingly, death is a non-relational possibility—it is a possibility that is uniquely my own.109 Likewise, when I turn toward my own limit—when I make death visible, I turn away from my everyday relations to others, and relate to myself alone. By relating to myself alone in Sein zum Tode, I am relating to myself authentically (eigentümlich) in a way that realizes my capacity for self-understanding: I confront my end (death) as the possibility that is uniquely my own. Without coming to terms with my own end—death—I live in-authentically. Not having realized its capacity for selfunderstanding, Dasein has not individualized itself and exists relationally in the They as Das Man. For Heidegger in Being and Time, achieving an authentic self-relation is an “impassioned freedom towards death.”110 By realizing my capacity for autonomous self-relation, the human being is freed from Das Man. Division II of Being and Time aims to bring to light the possibilities of Dasein as regards authenticity and totality—each of which are intimately related. But how can Dasein be an autonomous totality? If death is that which completes Dasein, and death is never actualized, then Dasein can never be a complete totality. Rather, it appears that Dasein is always

Emerging Recognition of the Significance of Indeterminacy  29 incomplete. In this way, we see that Heidegger—like his predecessors—links autonomy with indeterminacy. As Heidegger writes, “in Dasein there is undeniably a constant ‘lack of totality’ which finds an end with death.”111 Ordinarily, Dasein does not realize that it is incomplete, for in its ordinary Alltäglickhkeit, the human being aims to realize his possibilities—to bring them to completion. But in Being-towards-the-end (Sein Zum Ende), Dasein realizes the fundamental impossibility of realizing one’s limit—death—and thereby realizes the incompleteness of Dasein. In facing one’s own death, one understands oneself as incomplete—not all possibilities can be realized. In autonomous self-relation, Dasein realizes what it cannot be in its ordinary life; it realizes it’s incompleteness, its lack of totality. For Dasein to be free means for Dasein to understand its own inherent incompleteness. Of course, as with all things Heidegger, this is just the beginning of the story. Joseph P. Carter, in his brilliant essay Heidegger’s Sein Zum Tode as Radicalization of Aristotle’s Definition of Kinesis (κίνησις), illuminates being toward death by means of Aristotle’s definition of motion. Heidegger is clear that motion is the guiding thread for the explication of the Being of being-there of the human being.112 Motion is a way of being-there, interpreted as being-at-work. Heidegger does not read Dasein’s authentic activity as ἐντελέχεια, for Heidegger glosses ἐντελέχεια as “being at the end,” which Dasein cannot achieve, since once it is at its end (death), it can no longer be. Instead, Dasein is at work or actual in the traditional Greek terminology (ἐνέργεια) as a motion (κίνησις). For Aristotle, the motion is the actuality (being at work) of potential (δύναμις) as potential. For example, while a house is undergoing construction, the house does not yet exist actually— rather the house exists in the material only potentially, for the potential is coming to be actual. Motion is the reality of the coming to be of the actuality. In motion, the end of the motion exists outside the motion, for once the end comes to be, the motion ceases, e.g., once the house exists, the house building ceases to be. Usually when we speak of the realization of a possibility, we mean that the possibility no longer only exists as possibility, but also exists as actual. In the case of motion, the house is present only potentially—the potential to be a house is realized in motion—not the house itself. Accordingly, what is fully realized in the case of motion is the possibility as possibility, not the possibility as a separate actuality. Although Heidegger strips Dasein of teleology, if Dasein exists as motion, then Dasein only exists dynamically (namely potentially—δύναμις). Dasein only exists as long as the end, death, has not come to be—Dasein exists as that which moves toward death. Unlike Aristotle who prioritizes actuality over potentiality, such that motion can only exist if there is a separate and complete actuality that is the subject of motion, for Heidegger there is only truth if Dasein exists. Thus, Heidegger reverses the order of priority: the actualization of any possibility for Dasein depends upon the un-actualized possibility to die.

30  Gregory S. Moss, C.U.H.K. But death is not just the end of Dasein; rather, “death is a way to be.”113 The end is not realized as an actuality in the way that the statue exists actually in the marble; death can only be realized as a possibility. In understanding one’s incompleteness in being-toward-death, one realizes death as possibility. Heidegger illustrates this with a famous example: while ripening, the fruit is unripeness. The not-yet is already included in the fruits own being; in no way is the not-yet a random determination, but rather is something constitutive. Analogously, Dasein is alwaysalready its not-yet, so long as it is.114 What is fully actualized at every moment of Dasein’s being is his incompleteness. Dasein is fully actualized or complete only insofar as Dasein is incomplete. Accordingly, Dasein is present as an autonomous totality only insofar as it is incomplete. Just as the fruit is not fully ripe, Dasein is not dead, but death is still present and realized as not-yet,115 as the possibility that is actualized as a possibility at every moment of Dasein’s existence. In authenticity, Dasein runs ahead to its future (in anticipation) in its present. Authentic Dasein recognizes the pervasive presence of the not-yet, and in this way, makes death (its unsurpassable horizon) visibly present. Because possibilities are always only possibilities for Dasein, by making death present, the possibility of Death is realized or “set free” as a possibility for the first time. In the authentic way of being, Dasein realizes itself as motion: death (a potential of Dasein) is made actual as potential. Since death constitutes the limit of Dasein, in authentic life Dasein is present as a whole. Indeed, death is not only a way to be, but it is the way for Dasein to be whole: it is its potentiality-for-being-a-whole.116 Since the limit of Dasein is death, and death makes Dasein whole, then authentic Dasein understands its whole autonomous Being as always existing in potential. Authentic Dasein realizes its potential to die—but only as potential. In contrast, inauthentic Dasein imagines death as only present as the absence of Dasein. Inauthentic Dasein does not recognize the possibility that both Dasein can be and death can be actual. It fails to realize its death existentially, namely that potential to die is the always actualized potentiality in virtue of which Dasein has any possibility at all. Indeed, because that which is finite has its non-being within itself, and Dasein has its own non-being (death) within itself, Dasein is finite. Dasein’s limit, death, is not outside of Dasein or other to Dasein, as inauthentic life might imagine, but is actual within Dasein as its individuating potential to cease to be. Dasein’s sein is Seinkönnen.117 In sum, Dasein is given as a self-relating and autonomous whole insofar as it maintains its potential to die. Following the tradition of German philosophy, Heidegger recognizes that what exists by itself is fundamentally indeterminate. By actively maintaining its potential to die, Dasein realizes itself as fundamentally indeterminate. Heidegger’s inversion of Aristotle’s priority of actuality over potentiality

Emerging Recognition of the Significance of Indeterminacy  31 also engenders an inversion of the priority of the determinate over the indeterminate. Since potentiality-for-being-a-whole is prioritized over the actuality of any possibility, and potentiality as potentiality is indeterminate, Dasein’s autonomous self-relation is fundamentally indeterminate. Heidegger explicitly acknowledges this positive significance of indeterminacy in his concept of the certainty of death, conscience, and anxiety. With the certainty of death comes “the indefiniteness of its ‘when.’ ” “Everyday Being-towards-death evades this indefiniteness of certain death by conferring definiteness upon it.” Death is a possibility which is “certain and at the same time indefinite.”118 Dasein ought to “cultivate the indefiniteness of the certainty” of death.119 Conscience is that which calls Dasein to confront its death. Although the direction of conscience is determinate (back to oneself), conscience remains “obscure” and “indefinite.” Its content is indefinite and it calls “in the mode of keeping silent.” Not only is the self to whom the appeal is made indefinite, but the caller is also indefinite.120 This indeterminacy is not merely negative, but is a positive phenomenon: The particular indefiniteness of the caller and the impossibility of making more definite what this caller is are not just nothing; they are distinctive for it in a positive way.121 Anxiety is that attitude in which one can confront the indeterminate character of one’s own being.122 Anxiety brings us face to face with our death. Following Kierkegaard, Heidegger distinguishes anxiety from fear. Unlike fear, anxiety does not have a definite object. Rather, anxiety is of the indefinite: Therefore, I must point out that it [anxiety] is altogether different from fear and similar concepts that refer to something definite, whereas anxiety is freedom’s actuality as the possibility of possibility.123 Anxiety is the “dizziness of freedom” and “freedom’s possibility.”124 Anxiety is that state of mind constituted by motion, for it is the actuality of the possibility of possibility. We realize our capacity for dynamic existence only from a position of dynamic existence: we confront our indeterminate being (death) only from an indeterminate position (anxiety). In the varieties of existentialist thought human autonomy is a nondiscursive locus of indeterminacy. In Being and Nothingness, Jean Paul Sartre further develops Heidegger’s concept of freedom as preserving potentiality in his privileging of existence over essence. Sartre argues that self-consciousness must ultimately be conceived as fundamentally non-positional. Regarding Husserl, he writes that although with Husserl we grasp the eidetic structure of consciousness, “if in consciousness its existence must precede its essence, then both Descartes and Husserl have committed an error.”125 Both Descartes and Husserl conceive of the Cogito, the I think, as a determinate essence, which provides an absolute position from which knowing may be attained. If self-consciousness were positional, then an infinite regress would

32  Gregory S. Moss, C.U.H.K. ensue, in which every consciousness of an object would require another consciousness*, ad infinitum. Thus, there must be an “immediate non-cognitive relation of the self to itself.” Insofar as consciousness requires self-consciousness, every positional consciousness is also a non-positional consciousness of itself. Accordingly, self-consciousness cannot be construed in terms of knowledge (as it appears in Descartes, Hegel, and Husserl). Rather, non-reflective consciousness makes reflective consciousness possible, or what is the same: indeterminate self-relation makes possible all determinate self-relations.126 Sartre’s reflections on the condition of reflective self-consciousness illuminate a transformation in the concept of freedom. Rather than attempt to integrate the indeterminacy of the autonomous self-relation, αὐτὸ καθ᾽ αὑτὸ, in terms of knowledge, Sartre abandons the primacy of knowledge, because knowledge is always positional, and self-awareness is not fundamentally positional.127 Self-consciousness is not determined by anything other than itself. Thus, it is self-determined: its determinate existence of consciousness does not have its origin outside of itself.128 Because self-consciousness is freedom, and freedom is fundamentally indeterminate, self-consciousness is that which makes itself determinate: “there is no difference between the being of man and his being free.” Indeed, “human freedom precedes essence in man and makes it possible.”129 In sum, Sartre transforms free self-consciousness into the condition for the possibility of positional being and knowing: a “logic of freedom” (in Hegel’s sense) would be a contradiction in terms. Human existence (understood as indeterminate self-relation) precedes determinate self-relation (essence). Because freedom is indeterminate, it is indefinable and unnamable. Sartre asks: given that it is indefinable and unnamable, is it also indescribable?: How then are we to describe an existence which perpetually makes itself and which refuses to be confined in a definition?130 Any attempt to constrain freedom by a definition would limit it to some positional existence, and would thereby fail to uncover the condition for positional existence. Non-positional freedom from which all positions arise cannot be described without positing it as something it is not: positional or determinate existence, what Sartre calls “essence.” But Sartre is keen to point out that this apparent indescribability points to the positive significance of indeterminacy and the proper way to describe freedom: freedom is nihilation. Freedom appears to knowing awareness as something other than what can be said of it: it is that which can always negate whatever form it is in or what is the same: the self-relation of freedom is to negate whatever determinate limitation or position it posits.131 As Sartre points out: “no limits to my freedom can be found except freedom itself.”132 Freedom is the permanent possibility to initiate a rupture with myself and with the world. Absolute freedom means that I can always separate myself from what I have posited myself to be. The crisis of definition

Emerging Recognition of the Significance of Indeterminacy  33 appears as the definition: freedom is the presence of the irreducible potential to transform oneself into a new determination. Freedom temporalizes itself, or gives itself a history, through its activity of rupturing with its present and past positions. It is through these acts of rupturing that I learn of my freedom.133 In these ways, for Sartre indeterminacy has a deeply positive significance. Following Hegel, Sartre recognizes that autonomy, indeterminacy, and negation are intimately bound. Autonomous self-relation (the being of man) is indeterminate, and is thereby nothing: “if man as an existentialist sees him is not definable, it is because to begin with he is nothing.”134 But this is nothing less than the constant potential to realize itself as something.135 Irrespective of the something into which it has determined itself, because it can always initiate a rupture from that identity and determine itself anew, freedom is the inexhaustible potential to determine—to initiate separation (κρίσις): the positive significance of freedom lies in its indeterminacy—its capacity to exceeds itself. Insofar as it exceeds itself, indeterminacy is conceived as ecstatic—in the sense of Ekstasis, that which is outside itself. Ecstasy, another cognate of indeterminacy that has its roots in medieval descriptions of religious experience, is a common way of delineating indeterminacy in thinkers as diverse as Jacobi, Schelling, Heidegger, Sartre, and Nishitani. Freedom is never actually everything it could be. To be free means to preserve one’s potentiality—albeit in a different sense than Heidegger. Following Heidegger, it certainly privileges potentiality over actuality. To use a Hegelian phrase, freedom is the negation of negation, although it is without absolute determinacy, for the potential to negate every negation is an irreducible indeterminacy whose inexhaustible potential to be determinate constitutes its positive significance. Although Hannah Arendt has her own conception of freedom that is inseparable from the concept of plurality, she understands action (in contrast with labor and work) as the capacity to begin something new—to do the unexpected. Action is the realization of freedom, and is rooted in natality. Accordingly, Arendt emphasizes an aspect of freedom neglected by Heidegger’s focus on death: birth. Freedom and its indeterminacy is inseparable from the concept of birth: each birth is a new beginning. Action is constituted by a startling unexpectedness: the fact that man is capable of action means that the unexpected can be expected from him, that he is able to perform what is infinitely improbable. And this again is possible only because each man is unique, so that with each birth something uniquely new comes into the world.136 Although for Arendt it is both the case that action might be a less-frequent occurrence than Sartre’s conception of free action, and without others (plurality) action is not meaningful, we can still see that in Arendt’s conception

34  Gregory S. Moss, C.U.H.K. of freedom indeterminacy is paramount. Moreover, indeterminacy here links our discourse not only with beginning in general, but also with concerns of practical reason—such as political beginnings and political philosophy more generally. Trish Glazebrook and Michael Goldsby, in their Climate Science, Indeterminacy, and Food Security in Sub-Saharan Africa, offer a glimpse of how the concept of indeterminacy plays a significant role in the intersection between politics, ethics, and science. They show how conceptions of indeterminacy (the mismatch between the model and the target system) in scientific inquiry have been misappropriated and employed in misleading ways to both deny the reality of climate science and cause severe harm to vulnerable peoples. Certainly, at least in the domain of climate science, political action in Arendt’s sense seems more urgent than ever. Todd May, in The Effability of the Normative, engages with Adriana Cavarero’s Politics of Voice, and argues that independently of semantic form, normativity is at risk of becoming uncritical dogma. May emphasizes that although normativity is not reduced to linguistic standards, it is certainly bound to them. Steven Crowell also contributes to the conversation on normativity, when he argues that Merleau-Ponty’s conception of intentionality has not given projects their due. Crowell, in contrast, turns to Heidegger’s discussion of Dasein in Being and Time, where Heidegger accounts for the temporal world-horizon as a normative space of meaning. Since for Sartre free indeterminate consciousness undergirds reason and all positional being, no sufficient reason can be given for why the human being chooses to determine herself the way that she has. Following Heidegger, Sartre calls this “abandonment”:137 the human being is left alone, for there are no sets of values to justify or excuse one’s behavior. Freedom is without reason; it is radically spontaneous. Here we might invoke Sartre’s famous example of the young man who is faced with the choice to fight the occupying force or stay home to care for his mother. In Sartre’s analysis, neither reason, nor anything else, can determine in advance what he ought to do. Rather, the choice is thoroughly indeterminate. Rather, he decides what he ought to do by doing it. As Anthony Bruno points out in Determinacy, Indeterminacy, and Contingency in German Idealism, Schelling’s philosophy, long underprivileged and misunderstood as a stage in the development of Hegel’s absolute idealism, anticipates this existentialist position. For Schelling, freedom is the incomprehensible base for the reality of things, which forms the radically contingent base from which all ethical judgments are made. Indeed, to put the issue in terms of indeterminacy, the fact that philosophy is faced with the choice between integrating indeterminacy into philosophical science or problematizing its (philosophically) scientific character is itself evidence of the radical contingency of freedom. Long before Sartre transformed existentialism into atheism, Kierkegaard recognized the paradox inherent in free self-relation. The human being is free, and freedom is the “relation that relates to itself.”138 In order to fully realize the limits of one’s freedom, one must recognize that one does not

Emerging Recognition of the Significance of Indeterminacy  35 make oneself free. The human being is a derived and established self-relation. Although Sartre recognizes that we are condemned to be free, namely that the absolute and free self-relation is not self-established, he does not recognize God as the source of our freedom, as Kierkegaard does. For Kierkegaard, as long as we desire to be our own ground, namely to be the source of our very freedom, we remain self-abandoned and in despair, the solution for which is faith in the power that establishes freedom. Nonetheless, like Sartre, Kierkegaard acknowledges that insofar as freedom is “grounded transparently in the power that established it,”139 it must affirm a paradox that transcends reason and the principles of the ethical life. Realizing the power that establishes freedom requires the leap of faith. As is well known, Kierkegaard illustrates this leap with the example of Abraham. Abraham believed that God would be true to his promise that Isaac’s descendants would inherit the kingdom of Canaan even if Isaac were sacrificed by Abraham. Abraham had faith that the impossible is possible: He believed on the strength of the absurd, for there could be no question of human calculation and it was indeed absurd that God who demanded this of him should in the next instant withdrawal the demand. [. . .] His faith was not that he should be happy sometime in the hereafter, but that he should find blessed happiness here in this world. God could give him a new Isaac, bring the sacrificial offer back to life. He believed in the strength of the absurd, for all human calculation had long since been suspended.140 Since the leap of faith can only be performed on the strength of the absurd, and the philosopher cannot accept the truth of the absurd, the philosopher cannot perform the movement of faith. Rather, the movement of faith can only be marveled at.141 Accordingly, the full realization of human freedom transcends what philosophy can accomplish. The philosopher cannot endorse a contradiction—and this is exactly what faith engenders: What I am offering is a paradox. Yet I by no means think that faith is therefore something inferior, on the contrary that it is the highest, at the same time believing it dishonest of philosophy to offer something else instead and to slight faith. Philosophy cannot and should not give an account of faith, but should understand itself and know just what it has to offer.142 Although faith engenders a contradiction, and is necessary for the full realization of freedom, philosophy can recognize the need for something that it cannot accomplish. Against Hegel, Kierkegaard argues that philosophy ought to attempt to complete faith by means of dialectical reason: the absurd cannot be cancelled or preserved in a continuous process of dialectics. Rather, freedom is constituted by its discontinuity in a series of ruptures:

36  Gregory S. Moss, C.U.H.K. from the aesthetic to the ethical, and the ethical to the religious. The paradox of indeterminacy requires a leap of faith on the strength of the absurd; there is no appropriating into science: “all that can save him is the absurd; and this he grasps as faith.”143 Abraham’s faith is that “the single individual is higher than the universal.”144 In his dedication to God, Abraham intends to sacrifice his son. In this “teleological suspension of the ethical,” the lover (from the position of ethical reasoning) resembles the criminal,145 though is (paradoxically) higher than the universal, and thereby not to be judged by it. Kierkegaard’s insight is that for the philosopher, the non-positional is not a position that can be occupied, for to do so would be absurd. Since it is faith that believes in a truth that transcends reason, the truth of the indeterminate can only be approached in faith. By taking a stand where no stand can be taken, faith leaps into the indeterminate—into the absurd. Realizing freedom does not just involve making choices for which no reason is sufficient, but making choices that violate fundamental principles of rational thinking. Rather than simply reject the indeterminate as that which is not, Kierkegaard makes the indeterminate into an object of faith. Rather than integrate the indeterminate into free science, these various approaches lead us, albeit in different ways, to a learned ignorance. Indeed, philosophy begins in wonder: we do not know, but we desire to know. Philosophy begins without determinate wisdom. By working through the problem of indeterminacy, philosophy can reach a point of educated ignorance, wherein the indeterminate remains unknown, but this unknowing comes at the end of the process of thinking—it is no longer a naïve unknowing. In our naïve wonder, we may not know what the indeterminate is, but we may believe that we can in principle know it. Through our elucidation of the structure of knowledge, our unknowing becomes learned: in virtue of knowing the determinacy of knowledge, we develop an insight into why it is that we do not know the indeterminate. In Free Thinking in Schelling’s Erlangen Lectures,  I show that in his Erlangen lectures from the 1820s, Schelling argues that free thinking consists in revealing the indeterminate in and through our inability to successfully define it. Just at the point where thinking seems to fail to capture the indeterminate is exactly the point where thinking succeeds in experiencing it. Rather than successfully capturing knowledge of the whole in a proposition or judgment, the indeterminate is experienced in our failure to successfully capture it. The free and indeterminate Absolute is what runs through and transcends all determinations, though this can only be made clear in and through our attempt to determine it. Philosophy returns again to Socratic ignorance. Philosophy becomes again a practice in self-knowledge: γνῶθι σεαυτόv. But this return to Socrates is dialectical. In the case of Socrates, there is some determinate knowledge which as a human being he does not possess. In the failure to know it, the indeterminate shows up as that which it is: an exceeding of limits; in particular, that which exceeds the limit of knowledge. Rather than expect complete knowledge of the indeterminate,

Emerging Recognition of the Significance of Indeterminacy  37 the indeterminate is too great to know, and is sublime. William Desmond further develops the richness of philosophical wonder in Overdeterminacy, Affirming Indeterminacy, and the Dearth of Ontological Astonishment. In such ontological astonishment, the indeterminate becomes the over-determinate: the indeterminate as excess, which stupefies in the revelation of the unexpected. As sublime, the indeterminate becomes not only a concern for aesthetics, but also philosophy of religion. When the indeterminate becomes sublime, wonder becomes awe before a power that exceeds the power of reason. Rather than fruitlessly devote oneself to knowing a transcendent determinate reality (such as the Platonic Forms) in relation to which one can only stand in a position of ignorance, philosophy might instead become a practice of cultivating a disposition of awe before an indeterminate Absolute. By drawing on Michel Henry’s “Barbarism,” Aaron Simmons teaches us in Confidence Without Certainty that it is in art, poetry, and ethics where we work out meaning. Given the rise of scientism—the view that science is the sole source of truth—our capacity to work out meaning in art, poetry, and ethics remains an existential problem and task for thinking and acting. From early German Romanticism to the more recent theological turn, we see the aesthetic and theological dimension of indeterminacy brought to the fore. Considering Romanticism—we find that it is grounded on the philosophical thesis that the indeterminate Absolute is inconceivable. Indeed, the intimate connection between the problematizing of Logo-centrism and aesthetic categories is no more evident than in early German Romanticism. Schlegel recognized that every conception of the non-position rendered it positional. Although for philosophy this is an impossible contradiction, for Poesie it is not impossible. Phillip Mitchell explores Keats’ organic universal in Against the Darkness: Beauty and Indeterminacy in John Williams’s ”Stoner,” and the total vision of the poet, who holds opposites in unity. Since the non-positional has no a priori determination, there is nothing for reason to know. Hence, it falls to genius to give representational form to the sublime, namely for genius to spontaneously create the rule—which he cannot produce by imitation. In his article Indeterminacy, Gadamer, and Jazz, Bruce Benson not only illustrates the spontaneity and interplay of determinacy and indeterminacy in the improvisation and conventions of jazz, but he also shows that taste, like judgment, is indeterminate, for one can just as little give a rule for the former as one can for the latter. Rather than imitate a rule, early German Romantic philosophers argue that the inevitability of positioning the non-positional indicates instead that the indeterminate can be presented by the genius in allegory or the presentation of the unpresentable, as Schlegel defines it.146 Indeed—one can trace the roots of the turn to aesthetic categories in Continental philosophy to the early Romantic turn away from a discursive conception of the indeterminate to a poetic and aesthetic relationship. To put it in Nietzschean terms, indeterminacy calls us to the image

38  Gregory S. Moss, C.U.H.K. of Socrates in the Phaedo: Socrates must learn how to make music. Or put in the terms of Heidegger’s late philosophy: poetry is the primordial speech of Being. Indeed, the need to appeal to aesthetic categories to think or speak the Absolute reflects the real possibility that the indeterminate might simply be discursively inconceivable. In Jean Luc Marion’s theological return, St. Anselm’s famous definition of God as that than which nothing greater can be conceived is revived: the greatness or sublimity of God is inseparable from his inconceivability. That than which nothing greater can be conceived is the inconceivable itself: “God begin where the concept ends.”147 Accordingly, the theological turn is also a return, a recollection and re-appropriation of the ancient ἄπειρον: it is nothing less than a call to rethink the history, legacy, and determinacy of indeterminacy in philosophical history.

Chapter Summaries148 Part I In the first chapter, William Desmond draws a distinction between the indeterminate as the indefinite and an affirmative, richer sense of the indeterminate. While Hegel and others have primarily focused on the former sense of indeterminacy, the latter sense, Desmond argues, plays a positive role in making possible determinate knowledge and understanding. He goes on to discuss the link between the positive sense of indeterminacy and the porosity of being, the recognition of which, he argues, is integral to the capacity to experience what he calls ontological astonishment. Anthony Bruno (chapter 2) and Greg Moss (chapter 3) bring into focus the emphasis Schelling places on the significance of indeterminacy. Bruno addresses debates in German idealism that arise in response to the modal shift in logic, proposed by Kant, from a logic of thinking to a logic of experience. With the Kantian logic of experience, Bruno argues, arises a problem of radical contingency or what Bruno calls “rhapsodic determination” for logic. While Fichte and Hegel attempt to resolve the problem of contingency by constructing rational systems aimed at established the grounds for logic, Bruno shows how Schelling brings into view, in a proto-existentialist movement, the way in which the indeterminacy of the will undergirds rational system construction and, thereby, presents a limiting problem for the question of the value of system building itself. Moss presents an analysis of Schelling’s Erlangen Lectures, bringing into focus Schelling’s emphasis on preserving the significance of indeterminacy as integral to his re-envisioning of philosophy as ecstatic free thinking. As such, philosophical thinking extends beyond determinate knowing and involves what Moss characterizes as an ecstatic form of “intellectual intuition” that discloses, in a performative rather than conceptual manner, what cannot be comprehended through concepts.

Emerging Recognition of the Significance of Indeterminacy  39 In chapter 4, Indeterminacy, Modality, Dialectics: Hegel on the Possibility Not to Be, Nahum Brown investigates the theme of indeterminacy in relation to the question of unactualized possibilities. While the question of the role of unactualized possibility has been a significant theme in modal theories as diverse as that of Aristotle, Leibniz, and analytic modal logic, a specific version of the concept of unactualized possibility has preoccupied contemporary continental philosophers: the possibility not to be. This chapter focuses on Hegel’s analysis of the possibility not to be in the “Actuality” chapter of the Science of Logic and argues that Hegel’s modal theory is a precursor for continental modal theories such as Heidegger and Agamben’s. Brown argues that the possibility not to be plays a major role in Hegel’s modal theory and that emphasizing this motif in his work complicates the prevalent interpretation that Hegel is a philosopher principally concerned with actuality-primacy. Part II In chapter 5, Steven Crowell compares and contrasts the perspectives of Husserl, Merleau-Ponty, and Heidegger on the central phenomenological notion of horizon. For Husserl, Crowell argues, both the notion of “world” and of the horizon of objects play normatively regulative roles that, Husserl argues, make it possible for indeterminacies to become determinate and meaningful; however, in envisioning the task of phenomenological clarification as infinite, Husserl’s account problematizes the possibility of arriving at determinate meanings. By contrast, Merleau-Ponty’s understanding of horizons places greater emphasis on the importance of embodiment and the originary intentional relation of “motricity” in establishing the sense of things, and this leads Merleau-Ponty to allow for “ontological ambiguity” in the determination of sense, which, similar to Husserl, Crowell sees as an unresolved problem in Merleau-Ponty. Crowell suggests a solution to the problem of horizon that allows for the possibility of objective truths may be found in Heidegger’s description of specifically human comportment as involving both “trying to do” and “trying to be.” In chapter 6, Trish Glazebrook and Michael Goldsby highlight issues of gender, distributive, and intergenerational justice in relation to the disproportionate impacts climate change is having and will increasingly have on the global South. In doing so, they address the positive senses of uncertainty and indeterminacy in the context of rational, scientific inquiry and explain why climate deniers who rely on these terms to deter greenhouse gas mitigation policy have no rational basis for their position. Further, they argue that given the harm caused by inaction on climate change, especially to women, the poor, and future generations in the global South, the distortion of climate science by climate deniers not only lacks a rational basis, it is both unethical and inexcusable. In the following two chapters, Janet Donohoe and Robert Scott address the significance of horizons of indeterminacy in the later work of Husserl.

40  Gregory S. Moss, C.U.H.K. In chapter 7, Donohoe brings the problem of racism into dialogue with the ethical ramifications of the genetic method of phenomenology developed in Husserl’s work after 1919. In particular, she considers how horizons of indeterminacy play a role in what Husserl describes as a call for “questioning back” and for renewal and critique. She further considers how the ethical implications of genetic phenomenology may be usefully applied in reforming harmful inherited traditions and attitudes such as racism. In chapter 8, Scott brings into dialogue Husserlian phenomenology which, he argues, affirms the significance of indeterminacy, with Aristotelian metaphysics and ethical theory which are oriented towards complete determinacy. Drawing from the Husserlian notions of horizon and indeterminacy, Scott develops the idea of a horizon of indeterminacy at the limits of clear understanding which, he argues, is constitutive of a rational understanding of any object. Given the integral significance of indeterminacy to rational understanding, made evident by Husserlian phenomenology, Scott proceeds with a critical retrieval of Aristotle’s intellectual virtues according to which the highest intellectual virtue, philosophic wisdom, is reconceived as including recognition of indeterminacy and its practical significance as integral to the highest form of understanding. In chapter 9, Todd May engages critically with Levinas, Derrida, and others who, he argues, have tended to prioritize indeterminacy in the sense of ineffability as a means of grounding claims to normativity. May argues that while the ineffable may play an important role in establishing normative claims, the ineffable cannot be separated from effable semantic articulations that play an equally important role for establishing norms. Part III In Chapter 10, Bruce Benson reflects on the interplay of determinacy and indeterminacy in relation to hermeneutics, music, and genuine conversation. In a wide-ranging discussion that focuses on the link between the inextricable interplay of determinacy and indeterminacy in musical improvisation, Benson brings into dialogue aspects of Gadamer’s theory of hermeneutics with elements of John Cage’s and Roman Ingarden’s discussions of music. Both Nathan Eric Dickman (chapter 11) and J. Aaron Simmons (chapter 13) address the importance of recognizing the significance of indeterminacy as a vital counter to the contemporary impulse for easy, often misleading, answers. Addressing the bearing of indeterminacy on hermeneutic phenomenology, Dickman brings into focus the priority of “genuine questioning” in mediating between isolated experience and intersubjective understanding. In contrast to the contemporary impulse for quick and often over-simplified answers, Dickman valorizes the dialogical philosophical practices of genuine questioning and open-ended discussion which, he argues, are rooted in a shared appreciation of indeterminacy.

Emerging Recognition of the Significance of Indeterminacy  41 In chapter 12, Phillip Mitchell addresses the significance of indeterminacy from the perspective of a literary critic. Through a critical analysis of John William’s novel Stoner, a beautifully written work about an unheroic individual, Mitchell shows how a kind of indeterminacy arises in reading the novel through the tension between form and content. He further argues that the indeterminacy generated through the tension between the form and content of the novel provides “the necessary condition for the aesthetic experience” and, paradoxically, contributes to “completing the narrative.” J. Aaron Simmons takes up the question of the value of philosophy in an educational context increasingly oriented by scientism, STEM, market logic, and a desire for simple, certain answers. Through a critical engagement with works by Michel Henry and Kierkegaard, he argues that the unique value of philosophy consists in its capacity as a discipline to transmit to students the hermeneutic and existential value of open-ended questions and of humbly, yet confidently, embracing the indeterminacy of life. Part IV Both chapters 14 and 15 address the significance of indeterminacy in relation to the work of Dōgen Kigen. In chapter 12, Graham Priest and Filippo Casati draw connections between Martin Heidegger’s understanding of being as ineffable, and in that sense indeterminate, and Japanese Zen philosopher Dōgen Kigen’s understanding of ultimate reality as ineffable. Both Heidegger and Dōgen, they argue, adopt a dialetheist approach to engaging these basic concepts insofar as they affirm the contradictory practices of both speaking about what cannot be fully determined and remaining silent before it. In the following chapter, George Wrisley continues the discussion of Dōgen and considers ways in which the latter’s Mahayanan conception of the Bodhisattva ideal (which involves three focal indeterminacies, Wrisley notes) may be compared to and enriched by a Nietzschean conception of the higher type of individual. Qingjie James Wang (chapter 16) and Robert Neville (chapter 17) both consider how indeterminacy enters into Chinese philosophy. Wang focuses his analysis on the formative rather than normative character of Confucian ethics that arise from the Confucian embrace of the body (in contrast to traditional Western philosophy) as both intimately intertwined with the larger body of the cosmos and as the basis for the cultivation of both individual and collective virtue. In contrast to Western philosophy, which has tended to prioritize determinate knowledge and mind over body, Wang notes that the significance of the indeterminacy and porosity of the body have been upheld throughout the Confucian tradition insofar as the body is recognized as the primary link between the self, society, and the larger body of the cosmos. In chapter 17, Neville starts by comparing and contrasting Western and Chinese understandings of the significance of indeterminacy in relation to attempts to understand the ultimate source of being. Neville elucidates the

42  Gregory S. Moss, C.U.H.K. Daoist response to perennial cosmological questions by explaining the fundamentally indeterminate and creative dimension of the Dao by which the world emerges from nothing. Further, Neville clarifies important differences between Daoism and Confucianism while highlighting the ways in which both embrace a non-deterministic understanding of reality while emphasizing the importance of practical virtues. In the last chapter, Casati and Ricki Bliss pick up on a theme addressed by Neville in his discussion of Chinese philosophy. They consider the indeterminate (in the sense of ambiguous) meaning of the term nothing in the classic cosmological question of why there is something rather than nothing, with special focus on Martin Heidegger and Nishida Kitaro. They argue that recognizing the indeterminacy of the meaning of nothing can help clarify the sense in which this classic cosmological question remains significant for philosophical inquiry today.

Notes 1 Written by Gregory S. Moss 2 Barnes (2002, p. 18, [12B2]). 3 Aristotle (1999, 1029b13). 4 Ibid., p. 994a. 5 Plato (1903, 130b). 6 Barnes (2002, Heraclitus Fragment B51). 7 Plato (1925, 24c–25a). 8 Plato (1925, 25a). 9 Aristotle (1984, Book 3, Ch. 6). 10 Aristotle (1999, p. 174, 1048b), Sach’s translation. 11 Aristotle (1984, Book 3, Ch. 6). 12 Aristotle (1999, 982b). 13 Kant (1998a, p. 553); Kant (1998b, p. 652). 14 Kant (1998a, p. 555). 15 Ibid., p. 212, B106. 16 Ibid., p. 215, B111. 17 Hegel (2010, p. 101). Although Hegel eventually distinguishes negation from limitation within Chapter 2, Dasein, given our limited purpose and space restrictions, we will pass over the intricacies of this distinction. 18 Hegel (1969, p. 113). 19 Ibid., p. 157. 20 Ibid. 21 Hegel (2010, §21, p. 96). 22 Hegel (1969, p. 83). 23 Hegel (2010, p. 83 [21:97]). 24 Hegel (1969, p. 92). 25 Zhuangzi (2009, p. 54). 26 I am indebted to David Chai (CUHK) for offering this interpretation of the story. Although I intentionally put it aside, I recognize that this story also has import for practical philosophy. 27 Hegel (1969, p. 620). 28 Aristotle (1999, 1006b). 29 ἄπειρον ἄρα τὸ ἕν, εἰ μήτε ἀρχὴν μήτε τελευτὴν ἔχει. ἄπειρον. καὶ ἄνευ σχήματος ἄρα: οὔτε γὰρ [. . .] (Then the one, if it has neither beginning nor end, is unlimited.” “Yes, it is unlimited.” “And it is without form [. . .]”)

Emerging Recognition of the Significance of Indeterminacy  43 0 Eckhart (1981, “Sermon 6,” p. 187). 3 31 Ibid., “Sermon 52,” p. 201. 32 Plato (2005, 341c). 33 Plato (1925, 16c, 23c, 27b-27c). 34 Aristotle (1999, 209b 15). 35 ibid. p. 249, 1075a. 36 Kant (1998a, p. 554). 37 Ibid., p. 553. Kant amends the principle to state that it is a principle of thoroughgoing determination a priori, whereby entities acquire their predicates by having a share in the whole of all a priori predicates. Ibid., p. 554. 38 Ibid., p. 556. 39 Ibid., pp. 553–554. 40 Meiner (1998, B600, 652). 41 Kant (1998a, p. 555). 42 Ibid. 43 Ibid., pp. 556–557. 44 Ibid., p. 557. 45 Ibid., p. 557. 46 Ibid., p. 558. 47 Ibid., p. 554. 48 Ibid., p. 554 49 Heidegger (1967, p. 23). 50 Ibid., p. 26. 51 See Kant (1998a, pp. 253, 258, and 262). 52 Hegel (2005, p. 124). 53 See Hegel (2010, §21, p. 56). 54 Hegel (1969, p. 81) and Hegel (2005, p. 124). 55 Hegel (1969, p. 87). 56 Ibid., p. 92. 57 Ibid., p. 88. 58 Ibid., p. 152. 59 Ibid., p. 152. 60 Ibid., p. 81. 61 Of course, Kant’s construal of Nous is his own and is quite different from Aristotle’s definition. For Kant, Nous is intellectual intuition, in which the conceiving of the object would be sufficient to give the existence of the object in intuition. The Noumenon is that object which humans would know were they to be in possession of such a power. 62 Hegel (1969, pp. 164–165). 63 Ibid., pp. 389–390. 64 Ibid., p. 712. 65 Hegel (2005, p. 141); Hegel (1969, p. 157). 66 Hegel (1969, p. 603). 67 Husserl (1975, Ch. 3, Paragraph 27, p. 91). 68 Husserl (1974, p. 280). 69 Husserl (1975, p. 99). 70 Ibid., p. 98. 71 Husserl (1974, p. 279). 72 Husserl (2009). Here, I refer to the very conclusion of the piece, where Husserl advises against taking the beginning as “vorgegebenes” or “ueberliefertes.” 73 Husserl (1975, p. 98) 74 Ibid., p. 102. 75 Husserl (1974, p. 281). 76 Husserl (1975, p. 102). 77 Husserl (1965, p. 170).

44  Gregory S. Moss, C.U.H.K. 78 Husserl further categorizes horizons into visible and invisible horizons. 79 Geniusas (2012, p. 7). 80 This genetic analysis leads to new insights on the intersubjective constitution of subjectivity within a historical community. 81 Badiou, in Logic of Worlds, argues that the All could not exist. He arrives at the conclusion as a consequence of Russell’s paradox. Rather than one world, Badiou argues for a plurality of worlds. See Gabriel (2013, pp. xxv–xxvi), and Badiou (2009, p. 114). 82 Barnett (1998, p. 1). 83 Lyotard (1984, pp. 33–34). 84 Derrida (2016, p. 26). 85 Derrida (1981). 86 While the impact of Derrida and 20th-century French Continental philosophy on questions of indeterminacy have been widely acknowledged, our volume focuses more heavily on early German Continental philosophy and neglected figures and neglected global traditions, such as East Asian philosophy, the Kyoto School, and Schelling, whose philosophical engagement with indeterminacy have not been properly acknowledged. Nonetheless, indeterminacy in the French Continental tradition certainly deserves its own volume. 87 Nietzsche (1990, p. 45). 88 Ibid., p. 46. 89 Nietzsche, Truth and Lies in a Nonmoral Sense. Online at nietzsche.holtof. com/Nietzsche_various/on_truth_and_lies.htm 90 Ibid. 91 Nietzsche notes that he is afraid that we still believe in God because we still have faith in grammar. See Nietzsche (1990, p. 21). 92 Ibid., p. 177. 93 Ibid., p. 88. 94 Nishitani (1983, p. 150). 95 Ibid., p. 189. 96 Ibid., p. 122. 97 Ibid., p. 122. 98 Ibid., p. 74. 99 Ibid., p. 71. 100 Ibid., p. 73. 101 Ibid., p. 118. 102 Ibid. 103 Ibid., p. 140. 104 Ibid., p. 108. 105 Husserl (1970, p. 191). 106 Heidegger (1967, p. 171). 107 Ibid., p. 171. 108 Ibid., p. 307. 109 Ibid., p. 309. 110 Ibid., p. 311. 111 Ibid., p. 286. 112 Carter (2014, p. 474); GA (18:273); Cf. GA (33:172). 113 Heidegger (1967, p. 289). 114 Ibid., pp. 52 and 245. 115 Ibid., p. 289. 116 Ibid., p. 276. 117 Carter (2014, p. 496). 118 Heidegger (1967, p. 302). 119 Ibid., p. 301.

Emerging Recognition of the Significance of Indeterminacy  45 20 Ibid., pp. 317–319. 1 121 Ibid., p. 319. 122 Ibid., p. 310. 123 Kierkegaard (2000, p. 139). 124 Ibid., pp. 138–139. 125 Sartre (1992, p. 566). 126 Ibid., pp. 12–13. 127 Ibid., p. 17. 128 Ibid. p. 16. 129 Ibid., p. 60. 130 Ibid., p. 565. Although Sartre does not allow for us to speak of the freedom common to myself and the other, I am able to speak of my own consciousness (p. 566). 131 Ibid., p. 567. 132 Ibid., p. 567. 133 Ibid., p. 566. 134 Sartre (1989). 135 Ibid. 136 Arendt (1965, p. 21). 137 Sartre (1989). 138 Kierkegaard (2004, p. 43). 139 Ibid., p. 44. 140 Kierkegaard (1985, p. 65). 141 Ibid., pp. 66–67. 142 Ibid., p. 63. 143 Ibid., pp. 75–76. 144 Ibid., p. 84. 145 Kierkegaard (2009, p. 250). 146 Frank (2008, p. 208). 147 Marion (1999, p. 147). 148 Chapter summaries were prepared by Robert H. Scott.

Works Cited Aristotle. 1984. “Physics,” in The Complete Works of Aristotle, Vol. 1, edited by Jonathan Barnes. Princeton: Princeton University Press. ———. 1999. Metaphysics, translated by Joe Sachs. Sante Fe, NM: Green Lion Press. Arendt, Hannah. 1965. On Revolution. New York, NY: Viking Press, Revised second edition. Badiou, Alain. 2009. Logic of Worlds. Being and Event 2, translated by A. Toscano. New York, NY/London: Continuum. Barnes, Jonathan. 2002. Early Greek Philosophy. London: Penguin. Barnett, Stuart. ed. 1998. Hegel After Derrida. New York, NY: Routledge. Carter, Joseph P. 2014. “Heidegger’s Sein Zum Tode as Radicalization of Aristotle’s Definition of Kinesis.” Epoche: A Journal for the History of Philosophy 19(2): 473–502. Derrida, Jaques. 1981. Positions, translated by Alan Bass. London: The Althone Press. ———. 2016. Of Grammatology, translated by Gayatri Chakravotry Spivak. Baltimore, MD: John Hopkins Press. Eckhart, Meister. 1981. The Essential Sermons, Commentaries, Treatises, and Defense, translated by Edmund Colledge, O. S. A. and Bernard McGinn. Mahwah, NJ: Paulist Press.

46  Gregory S. Moss, C.U.H.K. Frank, Manfred. 2008/1990. The Philosophical Foundations of Early German Romanticism, translated by Elizabeth Millan Zaibert. Albany, NY: SUNY; New York, NY: Penguin. Gabriel, Markus. 2013. Transcendental Ontology: Essays in German Idealism. New York, NY: Bloomsbury. Geniusas, Saulius. 2012. The Origins of the Horizon in Husserl’s Phenomenology. Contributions to Phenomenology, Vol. 67. Dordrecht: Springer. Hegel, Georg Wilhelm Friedrich. 1969. Science of Logic, translated by Arnold V. Miller. Amherst, NY: Humanities Books. ———. 2005. Hegel’s Logic, translated by William Wallace. Oxford: Oxford University Press. ———. 2010. Science of Logic, translated and edited by George Di Giovanni. Cambridge: Cambridge University Press. Heidegger, Martin. 1967. Being and Time, translated by John Macquarie and Edward Robinson. New York, NY: Harper Collins. Husserl, Edmund. 1965. Philosophy and the Crisis of the European Man, translated by Quentin Lauer. New York, NY: Harper and Rowe. ———. 1970. The Crisis of the European Sciences and Transcendental Phenomenology, translated by David Carr. Evanston, IL: Northwestern University Press.—— —. 1974. Formale und Transzendnetale Logik. Versuch einer Kritik der logischen Vernunft. Leiden: Martinus Nijhoff. ———. 1975. Ideas Pertaining to a Pure Phenomenology and to a Phenomenological Philosophy, translated by Boyce Gibson. New York, NY: Collier MacMillan. ———. 2009. Philosophie als Strenge Wissenschaft. Hamburg: Meiner Verlag. Kant, Immanuel. 1998a. Critique of Pure Reason, translated and edity by Paul Guyer and Allen Wood. Cambridge: Cambridge University Press. ———. 1998b. Kritik der reinen Vernunft. Hamburg: Meiner Verlag. Kierkegaard, Soren. 1985. Fear and Trembling, translated by Alastair Hanna. London: Penguin. ———. 2000. “The Concept of Anxiety,” in The Essential Kierkegaard, edited by Howard V. Hong and Edna H. Hong. Princeton: Princeton University Press. ———. 2004. Sickness Unto Death, translated by Alastair Hannay. London: Penguin. ———. 2009. Works of Love. New York, NY: Harper Perennial Modern Classics. Lyotard, Jean-François. 1984. The Postmodern Condition: A Report on Knowledge, translated by Geoff Bennington and Brian Massumi. Minneapolis: Minneapolis University Press. Marion, Jean-Luc. 1999. Is the Ontological Argument Ontological? in Cartesian Questions: Method and Metaphysics. Chicago: University of Chicago Press. Nietzsche, Friedrich. Truth and Lies in a Nonmoral Sense. Online at nietzsche. holtof.com/Nietzsche_various/on_truth_and_lies.htm ———. 1990. Twilight of the Idols and the Anti-Christ, translated by R. J. Hollingdale. London: Penguin. Nishitani, Keiji. 1983. Religion and Nothingness. Oakland: University of California Press. Plato. 1903. “Parmenides,” in Platonis Opera, Vol. II, edited by John Burnet. Oxford: Oxford University Press.

Emerging Recognition of the Significance of Indeterminacy  47 ———. 1925. Philebus in Plato in Twelve Volumes, Vol. 9, translated by Harold N. Fowler. Cambridge, MA: Harvard University Press; London: William Heinemann Ltd. ———. 2005. The Seventh Letter. In Plato: The Collected Dialogues, edited by Edith Hamilton and Huntington Cairns. Princeton: Princeton University Press. Sartre, Jean Paul. 1989. “Existentialism Is a Humanism,” in Existentialism from Dostoyevsky to Sartre, edited by Walter Kaufman. New York, NY: Meridian Publishing Company. Online at www.marxists.org/reference/archive/sartre/works/ exist/sartre.htm ———. 1992. Being and Nothingness, translated by Hazel E. Barnes. New York, NY: Washington Square Press. Zhuangzi. 2009. Zhuangzi: The Essential Writings: With Selections from Traditional Commentaries, translated by Brook Ziporyn. Indianapolis, IN: Hackett Publishing.

Part I

The Significance of Indeterminacy in German Idealism

1 Overdeterminancy, Affirming Indeterminacy, and the Dearth of Ontological Astonishment William Desmond

From Indeterminacy to Overdeterminacy Looking back to ancient metaphysics, the indeterminate was dominantly seen in the light of the absence of intelligibility. Intelligibility was seen in terms of determinability or determinacy which, so to say, keeps at bay the formlessness of the indeterminate (to apeiron). The epistemic process by which being is made intelligible involves the movement of thought from the indeterminate to determinacy, the former being left behind in the process. We begin with an indeterminate wondering, pass through a more definite questioning or inquiry, and end with a more or less determinate answer to a well-defined question. The exploration here is of a more complex process than a teleological process from indeterminacy to determination. Among other things we need to invoke the self-determining and more importantly the overdeterminate in a sense to be explained. The point here will be not to negate the indeterminate but to offer what one might claim is an affirmative sense of the indeterminate, no longer as a formlessness to be transcended but a “too muchness” in being that calls forth our ontological astonishment. In the thoughts offered below I want to connect relevant issues here with Hegel’s understanding of thinking as negativity, claiming in connection with this that a dearth of ontological astonishment marks this understanding. First, I offer an overview of how we might look at the determinate, the indeterminate, the self-determining and the overdeterminate, and how they are related. The question of the dearth of ontological astonishment will then be addressed. In everyday realism, we think that things and processes have a more or less fixed and univocal character, and that this constitutes their determinacy. Nevertheless, determinacy cannot be understood purely in itself, but refers us to the outcome of the process of determination, a process not itself just another determinate thing. We tend to separate the determinate outcome from the determining process, and so take what is there as composed of a collection of determinate things. Determinacy is bound up with the fact that things and processes do manifest themselves with an immanent articulation, but whether that immanent articulation can be expressed entirely in

52  William Desmond univocal terms is an important question. If we put the stress only on univocity, we can cover over the process by which the determinate comes to be. Equivocal, dialectical and metaxological1 considerations enter into a fuller account of determinacy. By contrast with determinacy, the notion of indeterminacy is invoked. This might seem to be essentially a privative notion, referring us to the absence of determinate characteristics, and so hard to distinguish from what is void. A more affirmative understanding refers us to the matrix out of which determinate beings become determinate. As a kind of pre-determinate matrix, this reveals determining power in enabling the determinate things that come to be. I want to suggest that this more affirmative sense makes us think of the idea of overdeterminacy. Void indeterminacy refers us to an indefiniteness that is only the absence of determination, rather than the more fertile matrix out of which determinacy can come to be. These two senses of the indeterminate are often mixed up. If overdeterminacy is presupposed by indeterminacy, our general tendency to oscillate between the indeterminate and the determinate is shown not to go far enough. If determinacy is often correlated with univocity, and indeterminacy with equivocity, we need further dialectical and metaxological resources to do full justice to what is at play. An important consideration here that we need to consider also is the notion of self-determinacy. This refers us to a process of determination in which the unfolding recurs to itself and hence enters into self-relation in the very unfolding itself. This is particularly evident in the case of the human being as self-determining. The notion cannot be fully understood without reference to the ideas of the indeterminate and the determinate. Frequently self-determination is seen as the determination of the indeterminate in which a process of selving comes to achieve a relationship to itself. The human being is the most evident example of this, and particularly in modernity the idea of self-determination has received central attention. Both self-­determinacy and determinacy refer back to something that cannot be described in the terms of self-determination or determination. This something other, I suggest again, is not just the indeterminate understood in the privative sense, but the more affirmative sense which is the overdeterminate. Self-determinacy comes to be out of sources that are not just selfdetermining. Our powers of self-determining are endowed powers. There is a receiving of self before there is an acting of self. This makes the process of selving porous to sources of otherness that exceeds selving. The matter can be illuminated in a number of ways but I will refer to the speculative logic of Hegel where the triad, the indeterminate, the determinate, and the self-determining (sometimes via reciprocal determination) governs process as ultimately mediating itself in a self-becoming in which the other to the selving is the (self-)othering of the selving itself. The overdeterminacy does not enter systematically into the articulation of this understanding. By contrast, I will suggest that the overdeterminacy, as the affirmative sense of

Overdeterminancy, Affirming Indeterminacy  53 the indeterminate and not as the negative sense of the indefinite, refers us to the enabling matrix that makes possible determinacy and self-determination. The overdeterminacy has an excess more than all determinations, as well as more than what we can subject to self-determination. There is a “toomuchness” that has a primordial givenness that enables determinacy, that companions self-determination, yet also exceeds or outlives these. It is not to be equated with overdetermination understood as necessitation by an excess of determining causes. It allows rather the possibility of the open space of the indeterminate, and hence is not hyperbolic determinism, but hyperbolic to determinism in enabling the endowment of freedom. If Hegel’s dialectic tends to be defined by the triad of the indeterminate, the determinate and the self-determining, metaxology exceeds this triad in the direction of remaining true to the inexhaustible overdeterminacy. This inexhaustible overdeterminacy is multiply incarnated, for instance, in great artworks, or persons, or communities.

Ontological Astonishment and Hegelian Negativity To return to the widely present view throughout the philosophical tradition: to be intelligible is to be determinate; indeed, to be, properly speaking, is to be determinate. The status of determinacy and determination is at issue here. One might ask: How comes the determinate to be determinate? How does it come about that being as determinate is determinable by thought and hence rendered intelligible? The issue of a becoming determinate is at stake, not just some entirely static sense of being. One can see Hegel’s connection of thinking with determinate negation, or more generally with subjectivity as self-relating negativity, as answering to such questions. What is simply given to be is not intelligible as such; it is a mere immediacy till rendered intelligible, either through its own becoming intelligible, or through being made intelligible by thinking. Thinking as negativity moves us from the simple givenness of the “to be” to the more determinately intelligible; but the former (the “to be”) is no more than an indeterminacy, and hence deficient in true intelligibility, until this further development, determination has been made by thinking as negativity. A further complication in Hegel’s view is that thinking as a process of negation is not only a determining; it is in process toward knowing itself as a process of self-determining. Hence, his more complex description: self-relating negativity. The operation of negation is not only a determination of what is other to the thinking, for it is the coming to itself of the thinking process. In that sense, the return of thinking to itself, in the process of determining what is other, is not just making determinate, it is self-determining. Hence, the determining power of thinking in negativity is inseparable from Hegel’s understanding of the meaning of freedom. But there is logic overall that governs the movement of thinking as negativity: thinking moves from indeterminacy to determination to self-determination.

54  William Desmond In all of this there is a dearth of ontological astonishment, some aspects of which I propose to explore. For instance, given being as a mere indeterminate immediacy can barely be said to be, and even less said to be intelligible till rendered so by determining thinking that mediates by negativity. Hence, being becomes the most indigent of the categories that is all but nothing, till thought understands that it has already passed over into becoming. I don’t want to rehearse the famous opening of Hegel’s Logic, but want to suggest, among other things, that Hegelian negativity, via a logic of self-determining thought, is born of and leads to a dearth of ontological astonishment. Instead of a sense of being as the marvel of the “too much,” we find rather an indigence of “all but nothing.” I think we need to distinguish between three different modalities of wonder relevant to the issue: first, a more primal ontological astonishment that seeds metaphysical mindfulness; second, a restless perplexity in which thinking seeks to transcends initial indeterminacy toward more and more determinate outcomes; third, more determinate curiosity in which the initiating openness of wonder is dispelled in a determinate solution to a determinate problem. Determining thought answers to a powerful curiosity that renders intelligible the given, rather than to a primal astonishment before the marvel of the “to be” as given—given with a fullness impossible to describe in the language of negativity, though indeed in a certain sense it is no thing. Heidegger, for instance, has a truer sense of this other nothing. My focus is less defending Heidegger as to suggest the need to grant something more than a logos of becoming and self-becoming—there is an event of “coming to be” that asks of us a different logos. It asks of us a different sense of being, a different sense of nothing—not the nothing defining a determinate process of becoming, or a determining nothing defining a self-becoming: a nothing in relation to which a coming to be arises—a coming to be that is more primal than becoming. In a way, we can say nothing univocally direct about this nothing; rather we need to attend to how becoming and self-becoming presuppose this other sense of coming to be. A sense of this is communicated in the happening of a primal astonishment before the happening of the “to be.” I call this “overdeterminate” rather than just an indigent indeterminacy. In light of it, every process of determination and self-determination are secretly accompanied by what they cannot entirely accommodate on their own terms. This granting of the overdeterminacy of the “to be” has significance in relation to the dearth of ontological astonishment coming from understanding thinking as determinate negation, or self-relating negativity. It has a very important implication for the practise(s) of metaphysical thinking, especially one that tries to stay true to metaphysical wonder in the mode of primal astonishment. I will shortly say something about these three modalities of wonder, keeping in mind that wonder is not a univocal concept. It is not first a concept at all, but a happening, and as a happening it is plurivocal. The three modalities are internally related to each other, but they reveal a different stress in

Overdeterminancy, Affirming Indeterminacy  55 the unfolding of our porosity to being. If we do not properly attend to these different stresses, we can mistakenly think all wonder is subsumable into the curiosity that makes of all being an object of determinate cognition. This subsumption might consume curiosity, but it is the death of wonder. Wonder is not to be solely reconfigured as voracious curiosity that spends itself in ceaseless accumulation of determinate cognition. Equally, there is something other to thinking as self-relating negativity in wonder. In a way, one might say that we do not have a capacity for wonder; rather we are capacitated by wonder. Since this capacitation is not determined through ourselves alone, we alone cannot bring it to life or revive it. Wondering is not a power over which we exercise self-determination; it witnesses to a given porosity of being that endows us with the promise of mindfulness. If there is to be a vivification of the capacity, it is in coming home again to this porosity—and its capacitating of our powers. Ingredient in this homecoming is our capacity to know incapacitation.

Being Overdeterminate: Wonder as Astonishment Turning to the first modality of wonder as astonishment, I find it impossible to describe this astonishment in the language of negativity. There is a wonder preceding determinate and self-determining cognition that takes the form of a certain ontological astonishment. Wonder before the being there of being and beings is precipitated in this astonishment. This has not to do with a process of becoming this or that, but with porosity to the “that it is at all” of being. That being is, that beings have come to be at all, this is prior to their becoming this or that, prior to their self-becoming. In a certain sense, all human mindfulness is seeded in this astonishment. A caution: the word “wonder” strikes one today as a bit too subjectivized— it is seen as the “gosh” feeling, the “wow” experience to which we give vent before the surprising and the strange. One need not deny this gosh and wow but there is an ontological bite to original wonder, perhaps captured better in English with the word “astonishment.” In astonishment there is the stress of the emphatic: the unexpected is not anticipated to happen and yet it happens. When we say “The wonder of it is . . .” and refer to a happening, we are suggesting something beyond expectation—the surprising has communicated the emphatic. Being struck by astonishment has something of the blow of unpremeditated otherness in it. Extreme astonishment can seem even to deprive one of sensation. The blow of otherness stuns us, seems to stupefy us, as if inducing a kind of black-out. Many of the characteristics of astonishment—bewilderment, shock, consternation, deprivation of selfpossession, benumbing, being “stricken” by amazement—astound one even unto a kind of ontological stupor. All of this seems to be rife with a kind of negation—not our negation but our being negated. And yet it is more the affirmative “too-muchness” of the happening that is outlined in the event of astonishment. There is an

56  William Desmond intrusion of ontological frailty in the unpremeditated event of coming to be—it might not have come to be, it might not have been at all. And yet it is—surprising eventuation that hovers before us, floating above its own possible not being. It is hard for us to think on this boundary between being at all and possibly not being. Our porosity to the eventuation has the double character of itself happening as an opening, and being also a kind of “nothing.” This is not thinking as negativity but rather enables its possibility. This is evident in the fact that, in the opening of porosity, the rupture of surprise, while striking into us, takes us beyond ourselves: the self-transcending of thinking is possibilized. Astonishment is not just a subjective feeling. It is more like the seeding and first fertilization of the promise of “subjectivity” by an enigmatic communication to sleeping mindfulness from out of the intimate strangeness of being. We are moved into a between space where, in a sense, we go from our minds to the things; and yet there is no fixation of the difference of minding and things; our mindfulness wakes to itself by being woken up by the communication of being in its emphatic otherness. And so it is not quite that we go from our minds to the things, but more so that the things come to mind. Instead of thinking as negativity, already even before we more reflectively come to ourselves, there is the more primal opening in astonishment—an opening of which I would speak in terms of a certain porosity of being. In this porosity there is no fixed boundary between there and here, between outside and inside; there is a passage from what is into the awakening of mindfulness as, before any effort of its own self-determination, opened to what communicates to it from beyond itself. We do not open ourselves; being opened, we are as an opening. Astonishment awakens the porosity of mindfulness to being, in the communication of being to mindfulness, before mind comes to itself in more determinate form(s). In that respect also, it correlates with a more original “coming to be” prior to the formation of different processes of determinate becoming, and the more settled arrival of relatively determinate beings and processes. This is an important point in relation to the difference between wonder in the modalities of astonishment and curiosity. It is hard to think this more original porosity of astonishment, for all thinking already presupposes it as having happened. All determinate knowing proceeds from it, but it is not yet determinate knowing; nevertheless some sense of it can be communicated. I behold the majestic tree and murmur: “This is astonishing!” I am not projecting my feeling; I am being awakened by the tree, and am awakening to myself, in a more primal porosity, where the striking otherness of this blossoming presence has found its way into the intimate recesses of my now roused and receiving attendance. This astonishment is not a vector of intentionality that goes from subject to object; it is a porosity prior to intentionality, and hence refers us back to a patience of being more primal than any cognitive endeavour to be. Porosity might seem like negativity in that it cannot be reduced to this or that determination, and allows dynamism and passage. It is not thinking as negation but rather

Overdeterminancy, Affirming Indeterminacy  57 a mindful passio essendi prior to and presupposed by every conatus essendi of the mind desiring to understand this or that. First we do not desire to understand. Rather we are awoken or become awake in a not yet determinate minding that is not full with itself but filled with an openness to what is beyond itself—filled with openness, if that is permissible to say, for such a porosity looks like nothing determinate and, hence, seems almost nothing, even entirely empty. Being filled with openness and yet being empty: yet this is what makes possible all our determinate relations to determinate beings and processes, whether these relations be knowing ones or unknowing. Thinking understood primarily as negativity does not have enough of this porous patience, even though its endeavour to know ultimately derives from it. One might object that the desire to know is a drive to determination, a drive that when it comes to know itself becomes also more self-determining. It is a well rehearsed theme that philosophy begins in wonder and Aristotle is often cited: “All men desire to know” (Metaphysics, 982b11ff.). Aristotle sees the connection of marveling and astonishment when he reminds us of the affiliation of myth and metaphysics, and also the delight in the senses. Nevertheless, the desire to know is understood essentially as a drive to determinate intelligibility, which on being attained dissolves the initial wonder launching the quest. The end of Aristotle’s wonder is a determinate logos of a determinate somewhat, a tode ti. This end is the dissolution of wonder, not its deepening or refreshing. Significantly, Aristotle invokes geometry to illustrate the teleological thrust of the desire to know (Metaphysics, 983a13ff.). I take geometry here as representative of determinate cognition whose eureka solves the problem but also surpasses the wonder. I think the issue is better put in Plato’s Theaetetus (155d3–4), where thaumazein is named as the pathos of the philosopher. Pathos: there is a patience, a primal receptivity. This is not the self-activating knowing such as we have come to expect from Kant and his successors in German idealism, as well as in varieties of the constructivist epistemology we find in different contemporary inheritors of this Kantian stress. There is a pathos more primal than activity, a patience of the soul before any self-activity. One could say: there is no going beyond ourselves, no activation of our self-surpassing powers of transcending, without this more primal patience. I stress this since in modernity patience has often been relegated to a servile passivity supposedly beneath the high dignity of human power as self-activating, as self-determining. The truth is that no one can self-activate themselves into wonder. It comes or it does not come. We are struck into wonder. “Being struck” is beyond our self-determination. We cannot “project” ourselves into “being struck.” It comes to us from beyond ourselves. It does not come in the spiky oppugnancy of a hostile estrangement, though the hateful can strike one. It comes in the communication of an intimate strangeness that makes us porous to what before us is enigmatic and mysterious. Helpful here might be a brief comparison between first astonishment and curiosity (the third modality of wonder, to which I return). Curiosity is more

58  William Desmond to be correlated with a determinate cognition of a determinate somewhat (tode ti) or “object.” By contrast, in astonishment it is not that an “object” as other simply seizes us, making us passive while it is actively dominating. What is received, as we undergo it in “beholding from,” cannot be thus objectified. What seizes us is the offer of being beyond all objectification, and the call of truthfulness to being. This is not first either subjective or objective, but transsubjective and transobjective. “Trans”: we witness a crossing between “subject” and “object,” and an intermedium of their interplay that is more primordial than any determinable intermediation between the two. The happening of this “being-between,” in the occurrence of “beholding from,” reveals a porosity beyond subjectification and objectification and we are beholden to what eventuates in this between, making us answerable to its truth in our own being truthful. In the intimate strangeness of the porosity an excess of being flows, and overflows toward one. This is astonishing, not because initially we make no sense of it, but simply because the surprise of being’s being there at all is there at all. If there is something childlike about such a beginning, this does not mean it is merely childish. The childlike opening is our finding ourselves astonished already in the porosity of being. We do not produce astonishment; astonishment opens us in the first instance, and there is joy in the light. The child lives this primal and elemental opening; hence wonder is often noted as more characteristic of earlier stages of life. Thus too, as has been also noted, children have a spontaneous tendency to ask the “big questions.” First astonishment is more intimate with the primal porosity that constitutes the human being as metaphysically opened from the outset. The later developments of curiosity and sophisticated scientific knowing are seeded in the primal porosity but what its grant enables we too quickly take for granted. Then, alas, this maternal porosity can be long forgotten when the project of science comes more fully on the scene. When the child points to the night sky and murmurs—“Look, the moon!”—the astonishing has won its way into its heart. Later, the astonished child is recessed, even driven underground, in the curious project of (say) space exploration which lifts off the earth on the technical constructs of determinate cognition. The child is not only father to the man, but the man is the shield of time that shelters, or denies, the idiotic child it was originally born as. If the child dies, the shield shelters nothing, and the man dies too—a self-guarding hollowness, and not the elemental porosity. The callous of a self-circling conatus essendi covers over the idiotic pathos of the exposed child. This more-primal porosity of first given minding is at the origin of all modalities of mind, but as intimate with the giving of the first opening, it can be passed over, covered over. It enables the passage of mindfulness but the endowed passing can be passed over, since we come to ourselves in this passing. First a happening, it is only subsequently gathered to itself in an express self-relation. In this being gathered to itself, there is the risk of a contraction of what the first opening communicates. The gathering concretizes us

Overdeterminancy, Affirming Indeterminacy  59 as determinate, and as thus ontologically concentrated, we can contract the opening of the porosity to just what we will grant as given. I will come back to this when dealing with perplexity and curiosity, and with modes of minding that are determinate and self-determining. Nevertheless, as coming to mind in astonishment the porosity happens, we do not produce it, we do not determine it, it communicates from beyond our self-determination. Prior to the more determinate and determining selving of mindfulness the porosity that is neither of self or other happens as the between space in which, and out of which, a variety of determinate and self-determining forms of minding come to be. These latter are derived, not original. What is more original is the between of porosity.

Being Indeterminate: Wonder as Perplexity I now turn to perplexity as a second modality of wondering. We pass from original overdetermination to a mingling of indeterminacy and determination. In some ways, perplexity shows more evidence of the work of negativity in it. We are apt to think of perplexity as signifying our being troubled with doubt or uncertainty, our being puzzled. The word “plexus” in perplexity suggests a plaiting, a twining, an entanglement. We find the sense of something involved, com-plex, interwoven, something intricate and difficult to unravel, perhaps so knotted we wonder where to start with trying to untangle it. Plagued by perplexity, as we sometimes put it, our thoughts seem to be tormented with some vexing matter we cannot comprehend. Not only is it difficult to understand, but we may find ourselves thinking: we do not know what to think. There is nothing of calm serenity in this modality of wonder—there is often anxiety, bewilderment, distress, trouble, and perturbation. Perplexity arises out of first astonishment. How so? An important element of first astonishment is the way original wonder does not so overtake us as to squash us as selving but comes to release us into our more evident being for ourselves, into mindful selving as promising of itself, and perhaps of more than itself. We are granted to come to be ourselves, freed also into our own self-becoming in the between. This is part of what I call “the erotics of selving.” The original “too-muchness” of being is not indeterminate, not determinate, but exceeds all determination. In the first instance, it is overdeterminate, but as such endows the promise of self-determining. If astonishment holds the promise of the agapeic, there is awakened in it the erotics of self-surpassing. In being thus awakened, we are as selving, and come to ourselves as enabled more fully to become ourselves. Perhaps here some more express sense of thinking as negation can come into the open. Being awakened in the primal astonishment, the “too-muchness” of given being can seem to oppress us.2 Given to be as ourselves, the intimation intrudes that we cannot be its full measure. Though we are not the measure of the “too-muchness,” we yet want to know it in full measure.

60  William Desmond In this disjunction troubled perplexity arises: we do not know, we would know, we know we do not know. We are stressed in the baffling difference between what we know is too much for us, and our intimately known desire to know just that “too-muchness.” Perplexity is born in the baffling difference wherein our mindfulness is torn between its desire to know and its intimate knowing that it does not know what is too much for it. To live with this baffling difference is not easy, and there is the inevitable urge to diminish its stress in seeking a knowing that reduces the “too-muchness” to proportions that allow us to appropriate its difference. We are then faced with the urge to develop the desire to know as our way of subjecting the given “too-muchness” to our measure, that is, to the proportionate measure of ourselves as knowers. Often, perplexity takes off in this direction, but not always, and the entire situation is always more equivocal, since wonder as perplexity is recurrently haunted by faces of otherness that are just so as disproportionate to the determinate measure of our determinative cognition. Perplexity is a modality of wondering that brings us more into the equivocity of being: the play of light and darkness, the chiaroscuro of things and ourselves; the dark light of unformed things and things forming, of ourselves formless and seeking form and being returned to formlessness, of all things enigmatic and intimating, of ourselves the most baffling of beings, at once shouting absurdly and absurdly singing. Perplexity is not the reverse of astonishment but our waking to the troubling equivocity of the “toomuchness,” given in the astonishment. The equivocity is shown on both the sides of self-being and other-being. One might say the equivocity of the perplexing “too-muchness” is both transobjective and transsubjective. There is too much to the thereness of what is there; there is too much to the intimacy of being waking up to itself as our selving. Other-beings and selvings come from formlessness beyond form, are themselves as forming and coming to form, and finally point beyond themselves and all finite form. The troubling equivocity can fill us with great foreboding in face of the mystery of life, and every human being knows something of its disconcertment and dismay. It can drive us to distraction, it can drive us mad. It is never too much to say that it is always and ever too much for us to say. Perplexity awakens a seeking for what is true in all significant art, in all intellectually honest philosophy, in all spiritually serious religion. Mostly, however, the seeking has no fancy names, as ordinary persons in accustomed community, mostly out of the limelight, seek to tread the way of truth (with a bow to Parmenides). The equivocity of this perplexity is in the doubleness of being both the dismaying destitution of not-knowing and the ignorance of a voracious desire to know. Perplexity is first-born from original astonishment, but we wake up to ourselves even before and beyond the second-born desire to know. As a modality of wondering, something about perplexity is more primitive than what we normally call the desire to know. For we have already passed through treasures and dispossessions to get to the quotidian awakening of

Overdeterminancy, Affirming Indeterminacy  61 what we more ordinarily call the desire to know. This more primitive perplexity takes shape in the archaeology of the selving that comes to be out of the original ontological porosity. Of course, as transobjective and transsubjective, this perplexity is not just a matter of selving alone with itself. As coming to awakening out of the porosity, it is already an equivocal way of “being with” what is other than selving—a “being with” that is ingredient in waking selving both to itself and what is other to itself. Perplexity as wondering, like the primal astonishment, is a way of being between the “too-muchness” of other-being and selving coming to wakefulness of both itself and what is other. In wondering as perplexity, given the equivocal play of light and darkness, we are closer to something like Plato’s condition of the Cave. In perplexity, however, we are not in the Cave as prisoners who do not know they are prisoners. These latter do not know perplexity as an awakening. Perhaps these prisoners, that is, we ourselves in this condition, do have a dull presentiment that not all is as it seems. There is presentiment in perplexity but the dullness has already been tenderized into the pain of not being able to take for granted what now more and more enigmatically presents itself as being opened for questioning. To be perplexed is to realize that one is held in check by something too much for one’s own power. The chiaroscuro of being shows the troubling face of the equivocal “too-muchness” which holds us in a kind of thrall. To be enthralled is to be under a spell, but some thralls stop us, stupefy us. The “too-muchness” bewilders, befuddles, bemuses, bewitches us. Perplexity can be nonplussed by the equivocity. Nonplussed, we may appear to be stupid, but there is a salutary stupefaction in the wondering of perplexity. In moments of more ontological porous mindfulness that break into perplexity, we know that there is light, and that there is an access of light in perplexity truthfully undergone. That light might be the Siamese twin of the darkness, and yet the twinned darkness does not make it any the less the light. It is we ourselves who are twinned: participants in the perplexity which both burdens and enlightens, double-headed between the burden of the mystery and the godsend of light that gives ontological uplift. There is also the following deep equivocity. This we can see with an extremity of perplexity when it takes on the shape of horror: ontological horror before the being there of being in its excess to our rational measure. In the Cave we can turn downward as well as upward. Perplexity can come over us in the feeling of being blocked from ascending into the light. We would find light, but we find ourselves darkened—darkened in the very seeking for light itself. Not the measure of the light, we are also not the measure of this darkness. We cannot go up; perplexed we find ourselves falling. We may not want to fall, but we still find ourselves falling.3 Thinking as negativity may claim it can counteract the falling into equivocity by its progressive determination of intelligibility. The Hegelian way of doubt (der Weg des Zweifels—notice the reference to the double) will

62  William Desmond overcome radical equivocity through its own self-accomplishing skepticism.4 In accomplishing itself, skepticism overcomes skepticism, gives up its vagrancy (Kant described the skeptic as a nomad), and comes home to itself, in and as absolute knowing. Here knowing no longer feels the need to go beyond itself; it is finally at home with itself, having absolved itself from all alienating otherness, for all otherness proves finally to be its own otherness. It even surpasses the desire for wisdom, as in previous philosophy, and becomes possession of actual science, Wissensschaft. Previous philosophy was always between ignorance and wisdom; now there is no such between, since everything is between knowing and itself, in the circle of its own selfdetermination. In Hegel, after the old metaphysics, and the new critique, we are offered the new speculative philosophy which in post-transcendental form offers the totality of categories, each allegedly justified beyond critique, because of having been radically critiqued by dialectic. This dialectical way is carried on the labor of the negative to a mediation of the equivocity, through the many determinate intelligibilities, all the way to fully self-determining knowing. While this triadic movement from indeterminate, through determination, to self-determination has a certain qualified truth, it is not fully true to the dimensions of the perplexity suggested above. For here too there is something that exceeds determination, something also not to be described in the language of self-determining thought. If the latter take themselves to be the absolute measure of what is at issue, they suffer from the same bewitchment of the equivocity which they ostensibly claim to rationally mediate. They are within the Cave but have redefined its immanence as the whole, and hence are in an even worse position than those prisoners who know and grant with raw pain that they are still perplexed prisoners. The perplexity of the Cave has been dialectically domesticated: the Cave now is no Cave, since all there is is (self-)determined as immanence at home with itself and beyond which there is nothing greater to be thought. Without perplexity we settle into a false home at whose hearth flickers (self-determining immanence itself as) its own counterfeit god.

Being Determinate: Wonder as Curiosity I will close with a few words about curiosity as a third modality of wondering, one in which the overdeterminacy of astonishment is too easily forgotten, one in which the perplexity that can live on in thinking as negation is further dulled, even unto the death of wonder. If to be is to be determinate, here to be is nothing if it is not determinate. Being is nothing but determinacy and to be exhausted in the totality of all determinations. The danger: hostility to ontological astonishment is twinned with the annihilation of the wonder of being itself. Of course, we cannot but be curious, given our inextirpable desire to know the world around us and ourselves. The devil is in the details, or

Overdeterminancy, Affirming Indeterminacy  63 God is, we say; and often we think of the curious person, in his or her desire to know, as giving careful attention just to the details of things. Such attention, we think, can sometimes be carried to excess; it can be addressed to unworthy objects; we inquire into things, but too minutely. There is healthy curiosity; there seems also to be an undue or too intrusive inquisitiveness in which we are curious about what does not properly concern us. Curiosity, in a good sense, finds things interesting and surprising; its desire to know is open to the novel and strange; in turning to what is curious in things, inquiry fastens on their interesting determinacy, often with the twist of the odd. Novelty is important for the curious mind: the queer, the peculiar, whatever arouses closer attention. We also talk of a curious argument—one marked by ingeniousness or excessive nicety or subtlety. Those who are collectors of curiosities search in out-of-the-way places for things or people out of the ordinary. Interestingly, inquisitiveness, whether in approved or unapproved senses, can lead to inquisitions, in which novelty itself is suspect. The inquisitor is particular about details because the details revealed are unapproved. There is a desire to know what one has no right to know; prying curiosity intrudes on what properly does not concern it. This double-edged character means that, qua wonder, curiosity is not a pure porosity to what is true. What we are in the idiotic recesses of our being infiltrates our manners of being curious. There can be something closer to the purer porosity, the reception of astonishment, the awakening of perplexity. There can also surge up a will to know marked by a conatus essendi that wills to overtake, subordinate, if not extirpate the porosity and patience that are more intimate to the idiotic, ontological heart of our being. I stress this doubleness again, since one might claim that in our time this second possibility has taken on such an all-pervasive life that it seems to have an irresistible power of its own, and not really to come to be out of the more original porosity at the origins of wonder as astonishment and perplexity. If perplexity is a first-born child of primal astonishment, curiosity is a second-born. If astonishment is overdetermined, if perplexity mixes the overdeterminate and indeterminate, curiosity dominantly stresses the determinate. Often we think of wonder in this third modality as confronting problems. This is understandable—the “It is!” of first astonishment turns into the “What is?” (indeed “What the hell is it?”) of perplexity, turning now into the sober “What is it?” of curiosity. With this last form of the question, we ask about the determinate being there of beings, or the determinate forms or structures or processes. We move from ontological astonishment before being toward ontic regard concerning beings, their properties, patterns of developments, determinate formations, and so on. It is essential to the becoming of our mindfulness that we move into curiosity. The overdeterminate is saturated with determinations, not an indefiniteness empty of determinacy. The question “What is it?” turns toward the given intricacy of

64  William Desmond this, that, and the other, and there can be something even reverent in this turning, for it too shares in our porosity to the astonishing givenness. We can marvel as these given intricacies, coming to admire, and even be in awe of such immanent richness. Curiosity releases the self-surpassing energy of our questing to know in the mode of determinate questions bearing on this richness. There is, however, a certain understanding of curiosity that turns the teleology of wonder into a movement from the indeterminate to the determinate, and thence from determination to determination, all the way to the totality of determinations that are held to exhaust the whole. If we connect Hegelian negativity with a teleological movement from indeterminacy, through determination to self-determining knowing, his understanding is not quite to be identified with the view that being is simply determinate. Nevertheless, he does share in a crucial aspect of this teleology: what seems mysterious in the initial indeterminacy is brought into the light of full intelligibility at the end of the unfolding, in which intelligibility is rendered determinable by knowing as self-determining. This is evident at the highest level of absolute spirit: art comes to an end when the enigma of the origin no longer retains anything secret; in the end, religion safeguards no divine mystery that ultimately is too much for the power of philosophical knowing.5 Hegel’s self-determination thus shares this crucial orientation with this understanding of the teleology of curiosity. This kind of curiosity negates the indeterminate, for this as such cannot be grasped, for only the determinate is thus graspable. Behind this grasping can operate a metaphysical ressentiment against anything in the ontological situation that exceeds its measure, a secret hatred of the overdeterminate. Equally, all perplexity troubled by the “too-muchness” tends to be deemed an oppressive equivocity and as such no longer to be abided. There is no abiding with the mystery of given being. There is to be nothing abiding about the mystery of given being. If we conceive the teleology of knowing thus, and claim that this is the one and only path to the end of true knowing, the end result must be the evacuation of spiritual seriousness not only in art and religion but also in philosophy. We then suffer not simply from a dearth but from the death of ontological astonishment. For there is no room now for thaumazein in the modality of agapeic astonishment or in the modality of erotic perplexity. Great art works, like religious reverence or awe, may offer us striking occasions of originating wonder—ontological admiration, appreciation of being. If such wonder is entirely impelled out of its initial hiddenness by determinative curiosity, the porosity is no longer kept open in philosophical mindfulness. Philosophy, lacking the initiative of originating wonder, must itself atrophy, its ontological astonishment or perplexity substituted for by the virtuosity of technical cleverness or the second-hand scholasticism of commentary on commentary. It becomes treasonous to the wiser patience of first astonishment.

Overdeterminancy, Affirming Indeterminacy  65

Notes 1   For more on metaxology, see Desmond (2008). 2  Burke connects astonishment with horror: “astonishment is that state of the soul, in which all its motions are suspended, with some degree of horror” Burke (2015, Part II, section 1). He is not wrong but he is not entirely right either, since in terms of the analysis I offer here there is, in what he says of astonishment, a kind of mingling of wonder as agapeic astonishment and as erotic perplexity. The sense of horror becomes more overt, I think, on the turn of wonder from first astonishment to perplexity. Some of these nuances will be evident from my analysis, itself less psychological and more ontological-metaphysical than Burke’s. 3  Think of the aporiai of thought as showing a lack of poros: we are unable to find a way across, are at an impasse. In the Theaetetus, Plato again and again stresses the philosopher’s suffering of the aporetic. The question is related to perplexity. In the end, perplexity is not dissolved, but it can be the anticipation of a new occasion of trying to understand. It can also be addressed by myth or likely stories. Univocal theories are not enough. The way of philosophical perplexity is wayless. There is a noplace that is the place of thought (Socrates is described as atopos, see Symposium, Plato 2005, 215a2, 221c2–d6). This noplace witnessed to the porosity of the soul. Concerning perplexity, one also thinks of Kant and metaphysics: there are questions we cannot avoid raising but cannot also answer; we must raise them, but we cannot put them to rest in a univocal science or theory. Perplexity here is not like pure reason. It reminds one of trying to rest but being unable to find a comfortable position; one keeps casting around for a better position but finally the perplexity does not get dispelled. With some thinkers, it can be the opposite: they try to get away from perplexity by a strategy: “on the one hand, this,” “on the other hand, that.” Indeed, is there not much of zigzag in Kant? Perplexity can remind one of a fever where we restlessly turn this way, that way. Of course, this can generate the idea that thinking is itself a kind of sickness, reflection a curse, as happens with the underground man of Dostoevsky, and here and there with Nietzsche. The barbarism of reflection (see Vico [1999]) makes reflection itself the barbarism of the mind. This is perplexity sickened with itself, not the first astonishment, nor the posthumous wonder, I will discuss at the end. 4  der Weg des Zweifels—notice the reference to the double—is also described as a Weg der Verzweiflung—a pathway of despair. See Hegel (1952, §78, p. 67) and (1977, §78, p. 49; Phenomenology of Spirit as a “self-accomplishing skepticism (sich vollbringende Skeptizismus).” 5  On this in connection with art in relation to the teleological movement from symbolic, through classical, to romantic art, see Desmond (2003a, chapter 3). In connection with religion, see Desmond (2003b, especially chapter 6 in relation to the idea of Creation): creation is for him a “representation” that does not get to the true understanding which is “Creation” as God’s own self-determination. Creation is not the hyperbole of radical origination (see Desmond [2008, chapter 12]), and neither is the world as created the eventuation of finite being as given to be as other to the divine. The stress is not on such radical “coming to be” but first on becoming, then on self-becoming, indeed the self-becoming of God, and this following the teleological movement from indeterminacy, determination to selfdetermination. Just as there is no sense of hyperbolic giving to be, there is no sense of the baffling nothing out of which finite being is said to be given to be; there is determination, negation as the negativity immanent in the self-circling whole.

66  William Desmond

Works Cited Burke, Edmund. 2015. An Philosophical Enquiry into the Sublime and Beautiful. Oxford: Oxford World Classics. Desmond, William. 2003a. Art, Origins, Otherness. New York, NY: SUNY Press. ———. 2003b. Hegel’s God—A Counterfeit Double? Hampshire: Aldershot. ———. 2008. God and the Between. London: Blackwell. Hegel, G. W. F. 1952. Phänomenologie des Geistes. Hamburg: Felix Meiner. ———. 1977. Phenomenology of Spirit, translated by Arnold V. Miller. Oxford: Clarendon Press. Plato. 2005. “Symposium,” in Plato: The Collected Dialogues, edited by Edith Hamilton and Huntington Cairns. Princeton: Princeton University Press, 526–575. Vico, Giambattista. 1999. New Science, translated by David Marsh. New York, NY: Penguin Classics.

2 Determinacy, Indeterminacy, and Contingency in German Idealism G. Anthony Bruno

German idealism stands out in the history of philosophy for its systematic ambitions, despite, or perhaps because of, which it enjoys renewed interest. Guided by the idea that a philosophical tradition is better identified by its motivating problems than by its characteristic theses, we find German idealism driven by still-pressing questions. Are we rationally entitled to certain metaphysical concepts? Are such concepts merely empty forms lacking actuality? Is nature’s thoroughgoing explanation in terms of such concepts consistent with the nature and goals of human freedom? The space defined by these questions inspires Fichte, Hegel, and Schelling to develop a unified method for their solution after Kant. Early fissures appear in this method, which, while they do not undo the idealists’ shared motives, spawn a host of diverging agendas in subsequent post-Kantian thought. Determinacy and indeterminacy offer useful ways of viewing these fissures, particularly when we first consider a modal feature of Kant’s critical turn. In the Critique of Pure Reason, Kant reorients logic from an undisciplined use of the “form of thinking in general,”1 which yields the endless controversies of rationalist metaphysics, to an analysis of “the form of a possible experience in general.”2 Trading the logic of thinking for the logic of experience discloses a certain modal peculiarity of a priori conditions of possible experience, such as space, time, and the categories of the understanding. While necessary for us, they are radically contingent insofar as they lack a knowable, absolute ground: they are anthropically necessary, yet are brute facts.3 The critical turn accordingly confronts us with the radical contingency of the logic of experience. Kant’s tolerance for radical contingency initiates two disputes that shape the course of German idealism. The first concerns the set of a priori conditions. With no absolute ground, this set’s determination is not fully rigorous, but is to some extent haphazard or rhapsodic. To rectify this, Fichte and Hegel develop methods for determining the system of a priori conditions. The second dispute concerns the purpose or value of this system. Schelling argues that there is no decisive answer to why there should be such a system: however we construct it, its value is contingent on our “wholly undetermined” capacity to will its construction.4 This argument exposes German

68  G. Anthony Bruno idealism to the open question of systematicity’s value, confronting reason itself with its inborn indeterminacy of purpose. I propose to explore the concepts of determinacy and indeterminacy in order to trace German idealism’s path through these disputes. In §1, I explain Fichte’s charge that Kant determines the categories rhapsodically and outline his genetic deduction of a priori conditions of experience from the I. In §2, I illustrate this deduction’s application of the principle of determinability in the Foundations of Natural Right. In §3, I explicate Hegel’s rhapsody charge against Fichte before sketching his speculative determination of the system of conditions in the Science of Logic. Finally, in §4, I reconstruct Schelling’s argument in Philosophical Investigations into the Essence of Human Freedom that a philosophical system’s value is indeterminate because it issues from an originally undecided act of will. I hope to show that while determinacy guides German idealism’s highest initial ambitions, indeterminacy emerges as perhaps its natural and unavoidable limitation.

§1 Prior to giving a transcendental deduction of our entitlement to the categories in the first Critique, Kant offers a metaphysical deduction of the categories’ origin in the understanding. Whereas a transcendental deduction answers the question quid juris regarding our right to possess and use a pure concept, a metaphysical deduction answers the question quid facti regarding the fact from which that concept’s possession arises.5 Answering the question of right blunts skepticism about whether we are justified in using pure concepts like causality by showing that they are a priori conditions of experience. Answering the question of fact fills the lacuna left by explaining our possession of concepts that are necessary for the possibility of experience in terms of such contingent mechanisms as divine implantation and customary conjunction. Kant’s answer to the question quid facti is a metaphysical deduction that aims to show that the categories coincide with the logical forms of judgment.6 The faculty for judging—the understanding—provides an appropriate origin from which to derive these pure concepts insofar as it allows Kant to proceed systematically rather than “rhapsodically from a haphazard search for pure concepts, of the completeness of which one could never be certain, since one would only infer it through induction, without reflecting that in this way one would never see why just these and not other concepts should inhabit the pure understanding.”7 The charge of rhapsody concerns the modal status of the origin from which one derives the categories in response to the question quid facti. A derivation is rhapsodic if this origin is haphazard or contingent, which it is if, like the whim of a god or the flux of custom, a real alternative is possible. It consequently provokes uncertainty regarding the “completeness”

Determinacy, Indeterminacy, and Contingency  69 of the set of categories. By investigating the understanding’s own forms of judgment, Kant intends his metaphysical deduction to secure a rigorous determination of this set. Despite championing the spirit of transcendental idealism, Fichte returns the charge of rhapsody to Kant in the Wissenschaftslehre Nova Methodo: “Kant proves his philosophy only by means of induction and not through deduction.”8 Fichte here is targeting Kant’s metaphysical deduction, for, shortly after, he says that while Kant’s “conclusions are the same” as the Wissenschaftslehre’s, the latter “connects them to something higher.”9 Given that Kant’s transcendental deduction concludes with our right to the categories as necessary conditions of experience—a thesis that Fichte shares10—and given that Kant’s clue to this conclusion is the metaphysical deduction of the categories from something higher—namely, the understanding—Fichte’s charge must be that the metaphysical deduction is inductive. Why might this be, and why would Kant’s deduction thereby be rhapsodic? Fichte interprets Kant as arguing that experience is explicable if we assume “the operation of this or that [category]” and he asserts that this Kantian argument can secure “only hypothetical validity.”11 A hidden premise motivating Fichte’s assertion is that nothing definitively warrants the antecedent of Kant’s argument, namely, that we can assume the operation of the specific categories derived in the metaphysical deduction. Fichte’s premise would be true if there is any doubt about precisely which categories we may assume to be in operation. Now, Kant’s metaphysical deduction takes as operative those categories which coincide with the forms of judgment taken over from traditional logic—an inheritance that Kant himself recognizes is radically contingent in §21 of the Transcendental Analytic: for the peculiarity of our understanding, that it is able to bring about the unity of apperception a priori only by means of the categories and only through precisely this kind and number of them, a further ground may be offered just as little as one can be offered for why we have precisely these and no other functions for judgment or for why space and time are the sole forms of our possible intuition.12 Absent some “further,” absolute ground, the forms of judgment from which the categories are derived are facts as brute as the spatio-temporal character of human sensibility. This contingency extends to the categories, given their alleged coincidence with these forms. Hence, Kant’s determination of the set of a priori conditions of experience of which the categories are members is rhapsodic or, as Fichte says, inductive.13 The rhapsody problem is what elicits Fichte’s complaint in the Nova Methodo that “Kant does not derive the laws of human thinking in a rigorously scientific manner,” which, he claims, is “precisely what the Wissenschaftslehre is supposed to do.”14 Similarly, this problem is what prompts

70  G. Anthony Bruno him in Attempt at a New Presentation of the Wissenschaftslehre to ask of Kant, “who does not derive the presumed laws of the intellect from the very nature of the intellect [. . .] how did you become aware that the laws of the intellect are precisely these laws of substantiality and causality?”15 In order to secure a rigourous answer to the question quid facti—one that non-­rhapsodically determines the categories’ precise kind and number— Fichte’s Wissenschaftslehre begins by positing the freedom of the I as the “single basic law” and “explanatory ground of experience.”16 Such an absolute ground offers an apt origin from which to then deduce the categories insofar as it avoids Kant’s “detour” through traditionally observed logical forms, whose relative contingency leaves the categories’ completeness uncertain.17 “Nevertheless,” Fichte adds, it remains merely a presupposition that this constitutes the necessary and fundamental law of reason as a whole, a law from which we can derive the entire system of our necessary representations. [. . .] A complete transcendental idealism has to demonstrate the truth of this presupposition by actually providing a derivation of this system of representations, and precisely this constitutes its proper task. It does this by proceeding as follows: It shows that what is first set up as a fundamental principle, and directly demonstrated in consciousness, is impossible unless something else occurs along with it, and that this something else is impossible unless a third thing takes place, and so on until the conditions of what was first exhibited are completely exhausted, and this latter is, with respect to its possibility, fully intelligible.18 The I is a presupposition because it is a “fundamental principle” and, hence, not derivable. Nevertheless, as an absolute ground, it contains the demonstration of its truth in the form of a “derivation” of the system of its own “conditions.” During his Jena period, Fichte divides the Wissenschaftslehre into two parts: intellectually intuiting the I as first principle, and deducing from it the conditions of its realization.19 With this methodological division, the precise number and kind of categories are derived from the absolute ground whose realization they make possible. In order to distinguish the derivation of the system of a priori conditions from Kant’s own metaphysical and transcendental deductions, Fichte labels his deduction “genetic.” ’20 The notion of genesis is meant to reflect the sense in which these conditions are determined by a sui generis activity of thought. Anticipating the I’s emerging conditions, Fichte says: On the one hand, spatial extension and subsistence will be ascribed to it, and in this respect it becomes a determinate body; on the other hand, temporal identity and duration will be ascribed to it, and in this respect it becomes a soul. It is, however, the task of philosophy to demonstrate this and to provide a genetic account of how the

Determinacy, Indeterminacy, and Contingency  71 I comes to think of itself in these ways. Accordingly, this is not something philosophy has to presuppose, but rather is part of what has to be derived.21 Although the Wissenschaftslehre begins with a presupposition, the truth of its beginning cannot be presupposed. On pain of rhapsody, it cannot presuppose the conditions of its first principle’s realization, which include space, time, and the categories. A “complete transcendental idealism” must accordingly prove the truth of its beginning by genetically deducing these conditions from the I.22

§2 Fichte adopts several names for the rule that is meant to guide the determination of the system of a priori conditions. In the Nova Methodo, he says that deduction overcomes presupposition by following the “law of reflective opposition.” At this stage of the text, he has argued that there is “no consciousness of the I without consciousness of the Not-I.”23 This indicates that adhering to the rule in question consists in determining a condition through its opposite. Another name for this rule is the “law of reflection,” according to which “determining” something occurs “by means of opposition.”24 Fichte eventually refers to the rule as the “principle of determinability.”25 As a method of deducing a priori conditions, determination through opposition has the advantage of ensuring that no conditions are introduced rhapsodically, e.g., by inductive appeal to observed forms of judgment inherited from traditional logic. It instead proceeds by (a) positing a condition, (b) discovering that it is unstable without its opposite, and (c) generating a successor condition to resolve their resulting tension.26 A detailed deduction occurs in Foundations of Natural Right, where Fichte gives a “genetic proof”27 of the external world, the body, and other minds, among other a priori conditions. I will illustrate the principle of determinability’s application in an early stage of this proof, in which the concept of freedom is discovered to be unstable without that of a sensible world, the resulting tension between which concepts generates the concept of other minds, which resolves it. Like other texts in the Jena period, Natural Right posits I-hood as the “exclusive condition of all philosophizing.” ‘I-hood’ denotes a rational being’s existence as free—as “an acting upon itself.”28 Fichte claims that certain “necessary actions [. . .] follow from the concept of the rational being” that condition its possibility,29 the first of which we can ascertain via the principle of determinability. We find that the concept of freedom is unstable without that of an opposing not-I: “This activity is constrained and bound, if not with respect to its form (i.e., that the activity occurs), then with respect to its content (i.e., that the activity, once it occurs in a particular case, proceeds in a certain way).”30 While the form of freedom is to act upon itself and is otherwise

72  G. Anthony Bruno shunted blindly, it is empty without some distinct “content” to guide its movement. In other words, that freedom occurs is a condition of philosophizing, whereas how it occurs owes in part to some not-I. Hence, if free activity is to be “an efficacy directed at objects,” then some “world-intuiting activity” must limit it.31 Freedom’s determinacy accordingly requires positing a sensible world in opposition to it.32 Fichte acknowledges that common sense already grants existence to the world without the aid of philosophy, but he asserts that this bestowal must be explained. Philosophy must “bracket” common sense in order to explain “why we can posit ourselves only as altering the form of things, but never the matter.”33 According to Fichte’s explanation, it is because we cannot determinately posit ourselves as free without an opposing not-I that we posit the independent matter of the world.34 A tension now emerges. At one stage, a subject limits the world insofar as she is free (a). However, at another stage, the world limits her insofar as she does not create the world (b). A “prior moment” precedes the world’s limitation by the subject, one in which she is limited by the world, “and so ad infinitum.”35 This regress, Fichte says, “can be cancelled only if it is assumed that the subject’s efficacy is synthetically unified with the object in one and the same moment.”36 In particular, the synthesis of the subject and what opposes it requires an event that “leave[s] the subject in full possession of its freedom to be self-determining.”37 Fichte locates this event in another’s summons, which regards “the subject’s being-determined as its being-determined to be self-determining [. . .] calling upon it to resolve to exercise its efficacy.”38 Your summons limits me by opposing my agency, yet limits itself by assuming and inviting the exercise of the same: as we might say, it imposes a normative rather than a factual limitation.39 It thereby unifies subject and object, resolving the tension of the subject’s limitation by the sensible world (c). In the causally more robust space of a social world, the summons’ final cause or “ultimate end” is that the addressee’s freedom “ought to exist.”40 You make room for my free activity in a way that a merely sensible world cannot. Fichte thus infers that the “undivided event” of “free reciprocal efficacy” between minds is an a priori condition of experience.41 The concept of another mind summoning me to action affords stable determinacy to the concepts of freedom and world. Fichte arrives at this intermediate deductive result via the principle of determinability. Without mentioning this principle, he states: “In this process of distinguishing through opposition, the subject acts in such a way that the concept of itself as a free being and the concept of the rational being outside it (as a free being like itself) are mutually determined and conditioned.”42 The instability of the concept of freedom, and the tension that it produces with the concept of a merely sensible world, generate the first in a series of a priori conditions whose totality Fichte gradually determines with a genetic deduction from a first principle.43

Determinacy, Indeterminacy, and Contingency  73

§3 Kant’s tolerance for radically contingent, i.e., absolutely groundless a priori conditions provokes Fichte’s charge of rhapsody and inspires him to deduce such conditions from the I. Deducing them via the principle of determinability is meant to remove their contingency. Nevertheless, this deduction rests on the contingency of its first principle, for while the I is meant to generate the necessary conditions of its own possibility, it is itself “merely a presupposition.” Although self-positing, the I as first principle is not a given datum or a finished fact, but an activity that we must resolve to perform, on pain of what Fichte calls “dogmatism,” by which he means Spinozism. We will see that, for Hegel, this presents a Kantian relapse that impedes the rigorous determination of the system of conditions. In the Encyclopedia Logic, Hegel hails Fichte for detecting the rhapsody problem: It remains the profound and enduring merit of Fichte’s philosophy to have reminded us that the thought-determinations [i.e., the categories] must be exhibited in their necessity, and that it is essential for them to be deduced.—Fichte’s philosophy ought to have had at least this effect upon the method of presenting a treatise on logic: that the thoughtdeterminations in general [. . .] are no longer just taken from observation and thus apprehended only empirically, but are deduced from thinking itself. If thinking has to be able to prove anything at all [. . .] then it must above all be capable of proving its very own peculiar content, and able to gain insight into the necessity of this content.44 Hegel holds that if the categories are “thought-determinations”—as required by an adequate answer to the question quid facti—they cannot be “empirically” grasped or “taken from observation” of traditional or customary rules, but rather must be “deduced from thinking itself.” This echoes Fichte’s charge of rhapsody against Kant and his subsequent demand that the categories be derived “from the very nature of the intellect.”45 Nevertheless, Hegel holds that logic, in order to be a rigorous “Science,” must begin from “total presuppositionlessness.” Lacking any further ground, a presupposition is radically contingent—merely “an arbitrary assurance.”46 As he says in the Science of Logic, that which “has a presupposition [. . .] takes its start from the contingent.”47 Hegel, accordingly, has reason to suspect a Kantian tolerance for contingency in Fichte’s first principle, which we saw is “merely a presupposition.” Moreover, in Concerning the Concept of the Wissenschaftslehre, Fichte says that his science “is not something that exists independently of us and without our help. On the contrary, it is something which can only be produced by the freedom of our mind, turned in a particular direction.”48 Inseparable from the Wissenschaftslehre is the contingency of our adopting its orienting standpoint.

74  G. Anthony Bruno The question, as Fichte says in the New Presentation, is whether we assign the “explanatory ground of experience” to the I and embrace “idealism” or assign it to the not-I and embrace “dogmatism.”49 Answering this question is philosophy’s first task. Answering it in favor of idealism is, moreover, its “first demand.”50 But this presupposes one’s “confidence in one’s own selfsufficiency and freedom.”51 Whether I affirm my freedom and posit the I is a radically contingent or brute fact. For Hegel, by contrast, presuppositions must “be given up when we enter into the Science.”52 Fichte’s violation of Hegel’s scientific criterion extends even further. In Faith and Knowledge, Hegel notes that the Fichtean I “is not absolute” insofar as it is “conditioned by something else,” namely, the very need to deduce its conditions of realization. The I’s “incompleteness” in this regard is what necessitates “the deduction of the world of sense.”53 Worse still, the sensible world offers dubious determinacy to the I’s free activity, for it “appears as an incomprehensible primitive determinateness,”54 i.e., as a brute fact. Fichte appears to presuppose the world, since, in order to posit an (incomplete) I, he must first have “abstracted from the alien other which is afterwards taken back again.” Indeed, that the I initially lacks a proof of its truth implies its limitation in advance by something other. Hegel accordingly describes Fichte’s I as a “mirror” that “receives the sense-world and posits it ideally within itself, only to give it back afterward just as it received it.”55 If this assessment is correct, then the first presupposition of the Wissenschaftslehre is saddled with a second. Such a starting point cannot help initiating a rhapsodic determination of a priori conditions.56 It, therefore, cannot “display the realm of thought philosophically, that is, in its own immanent activity or, what is the same, in its necessary development.”57 Beyond presuppositionlessness, Hegel is concerned for the very thesis that drives Fichte’s charge of rhapsody against Kant: reason’s absolute freedom.58 Fichte sees that a rigourous answer to the question quid facti must show that the origin of the conditions of experience does not exceed reason’s power of explanation. Eliminating externality in this respect is meant to enshrine the freedom of reason or I-hood. However, Hegel observes in the Encyclopedia that the I “does not genuinely appear as free, spontaneous activity [. . .] having been aroused only by a check from outside.”59 Hegel refers here to Fichte’s claim in Foundations of the Entire Wissenschaftslehre that “the ultimate ground of all reality for the I is an original interaction between the I and some other thing outside it,” which he calls a “check.”60 If Fichte presupposes not only the I as first principle, but also the world as a check on the I’s realization, then he entrenches the threat of radical contingency and so fails to overcome Kant’s restriction on reason’s capacity for self-explanation. Fichte might reply that the idea of a check is underdeveloped in the Foundations, as it does not signify the social world that is deduced in Natural Right, from which he may infer that Hegel’s attack is stalled at a dialectical stage in which the tension between opposing concepts has not yet been

Determinacy, Indeterminacy, and Contingency  75 resolved. But Hegel can retort that the very idea of the opposition between the I and the not-I is arbitrary because the I already abstracts from the not-I, “an incomprehensible primitive” or “bare assurance” imposed on the I. Whether subsequently analyzed as the subject’s sensible object or as its social other, the not-I’s initial opposition lacks deductive necessity. Consequently, determining the I through its opposition to the not-I, in accordance with the principle of determinability, is at best incomplete. However, rather than abandon Fichte’s method of determining a priori conditions through opposition, Hegel adopts a modified method according to which opposition is necessary insofar as it comes to originate in a condition. After stating his scientific criterion, Hegel gives a preliminary gloss of this new method: “the logical has three sides: (α) the side of abstraction or of the understanding, (β) the dialectical or negatively rational side, [and] (γ) the speculative or positively rational one.”61 Understanding initially captures Fichte’s awkward pivot on two original presuppositions: through “restricted abstraction,” understanding “stops short” at the opposition of brute or “fixed” determinacies. Dialectic then transforms Fichte’s notion of opposition from the arbitrary and external imposition on a condition to that condition’s own inner contradiction: as Hegel says, dialectic is “the genuine nature” of a condition, a nature that “is not restricted merely from the outside,” but “passes over, of itself, into its opposite”;62 a condition thus negates itself through the contradiction that it contains. Finally, speculation yields a “positive result” where Fichtean deduction secures only rhapsody: it produces “not simple, formal unity, but a unity of distinct determinations,” namely, a unity of a condition and its own inner opposition.63 I will trace this method through the first stage of the Science of Logic in order to illustrate Hegel’s systematization of Fichte’s response to the question quid facti.64 Hegel’s Logic takes being as the least arbitrary starting point. Since being is all encompassing, it entails no differentiating (hence, no potentially arbitrary) opposition. As Hegel intones: “Being, pure being—without further determination. In its indeterminate immediacy it is equal only to itself and also not unequal with respect to another; it has no difference within it, nor any outwardly.”65 Being avoids the fixed and mutually external abstractions of the understanding that saddle Fichte with two equiprimordial presuppositions. But being also initiates the negations of dialectic, for its indeterminacy leaves “as little” for thought as the concept of nothing. Indeed, Hegel says, “There is nothing to be intuited in it.”66 Given its “emptiness,” being passes “of itself” into its opposite. It negates itself as all-encompassing insofar as it harbours its own contradiction.67 This yields “distinct determinations” ripe for speculative unification. Although being and nothing are indistinguishable in their indeterminacy and so are, in some sense, “the same,” they do not constitute a simple, stable unity. Rather, their unity is the “movement” of “each immediately vanish[ing] in its opposite,”68 a movement that Hegel

76  G. Anthony Bruno calls “becoming.” Becoming is their positive result—the first in a long series of thought-determinations whose totality Hegel calls “the absolute idea.”69 By removing Fichtean presuppositions from the system of a priori conditions, Hegel provides a deeper solution to the rhapsody problem. Nevertheless, his solution exhibits yet a deeper contingency, for systematicity raises the question of its value, the indeterminacy of which, Schelling will argue, reveals reason’s insuperable limits.

§4 Hegel rejects haphazard or rhapsodic limits on reason’s power for explanation in order to demonstrate its absolute freedom. While this systematically articulates how reason’s essence must be, that reason exists at all is a separate matter, one that Schelling continually investigates through variations on the question, “Why is there something rather than nothing?”70 No appeal to reason can answer this question without raising it anew, for even if the rigourous determinacy of a priori conditions is entirely internal to a rational system, that system’s value and the purpose of its construction originate in a brute act of will.71 Turning to German idealism’s final shift, I will reconstruct Schelling’s argument in the Freedom essay that freedom is an undetermined capacity for willing. It is useful to begin by noting that the Freedom essay abandons two doctrines from Schelling’s earlier work. First, it rejects absolute knowledge. In Of the I As First Principle, Schelling had posited “an ultimate point of reality on which everything depends,” of which we have knowledge “through which alone all other knowledge is knowledge.” This point is the I, and this knowledge is intellectual intuition.72 While Schelling does not mention intellectual intuition in the Freedom essay, he rejects absolute knowledge when he chides those who lament that the ground of reality and knowledge is “incomprehensible” and “without understanding.”73 Second, the essay rejects absolute idealism, the view promulgated in Schelling’s identity philosophy, according to which reason has absolute knowledge of itself, while we finite knowers are “merely its organ.”74 In stark contrast, he now declares individual human freedom to be “the one and all of philosophy.”75 Now, in the essay, Schelling charges Fichte with “subjective idealism” for failing to show that “everything actual” has “freedom as its ground.”76 But what precisely is Fichte’s error, given that he explicitly posits the I’s freedom as absolute ground? First, it is the “arrogance” of holding that the absolute can be known and can thus be given “order and form.”77 Second, it is the “impetuosity” of lamenting that the “darkness” of a “will in which there is no understanding” should be “the root of understanding.”78 Both claims reprise Schelling’s earlier criticism of the doctrine of intellectual intuition.79 The second claim in particular inspires a decades-long critique of Hegel, casting freedom as “the incomprehensible base of reality in things.”80 What is Schelling’s argument for this claim?

Determinacy, Indeterminacy, and Contingency  77 Freedom is incomprehensible if it is ultimately undetermined by reason. Schelling supports the antecedent by claiming that freedom, regarded as “a wholly undetermined capacity to will one or the other of two contradictory opposites, without determining reasons but simply because it is willed, has in fact the original undecidedness of human being as idea in its favor.”81 The supporting claim is that human existence is always to be decided. Who I am to be remains an open question. Schelling immediately clarifies that the will’s indeterminacy in this respect does not refer to “individual actions,”82 which must have their determining reasons. Rather, that I commit to a system or way of life for which such reasons can appear as determining—e.g., the system of Kantian ethics—is an originally undecided act of will.83 Freedom so regarded is therefore a capacity to open a space in which certain reasons go unquestioned, yet is itself unfathomable—an “unground.”84 Schelling’s argument casts in unintended light Hegel’s claim that Science begins with the “resolve, which can also be viewed as arbitrary, of considering thinking as such.”85 If, as with any way of life, a philosophical system rests on an “arbitrary” resolve—driven, say, by a longing for Science—then no system is presuppositionless.86 A system presupposes a radically contingent valuation, that is, an act of freedom determined in advance by no principle of reason, but driven by a “yearning and desire” for understanding. Accordingly, while freedom expresses a “will in which there is no understanding,” it is not a mere drive, but a drive toward systematic intelligibility— what Schelling calls “a will of the understanding.”87 To be sure, Schelling speaks in the essay of freedom’s “inner necessity,” an “essence” that he defines as “fundamentally [one’s] own act.” However, if my essence is an “act”, as opposed to a principle of reason, then I am originally “an undecided being.”88 Freedom’s “inner necessity” accordingly consists in one’s ineluctable responsibility to resolve how to live, which decision is radically contingent. Contra Hegel, a system’s “innermost presupposition” is freedom, whose essence is indeterminacy or “absolute indifference.”89 As Schelling says in his 1827/28 Munich lectures: “Common ethical judgment therefore recognizes in every person—and to that extent in everything—a region in which there is no reason at all, but rather absolute freedom. [. . .] The unreason of eternity lies this close in every person, and they are horrified by it as it is brought to their consciousness.”90 If freedom is the indeterminate and therefore incomprehensible ground of reason, then a presuppositionless system is a contradiction in terms, for a system’s driving value—why it matters—issues from a brute act of will. Schelling’s defense of this claim revives a Kantian tolerance for radical contingency, thwarting the thesis for reason’s absolute autonomy that inspires Fichte’s and Hegel’s systematic constructions. It is precisely the original undecidedness of human freedom that, for Schelling, reveals reason’s inborn indeterminacy of purpose. At the heart of the German idealist tradition—between alleged subjective idealism and professed absolute idealism—emerges the proto-existentialist insight that, however we answer

78  G. Anthony Bruno perennial questions about metaphysical concepts, their actuality, and their explanatory relation to human freedom, we inescapably embody the further question of why such questions matter.

Notes 1 Kant (1998, A55/B79). 2 Ibid., A246/B303. 3 Allison (2006) argues that what distinguishes Kant’s “anthropological” turn from empiricism is the latter’s retention of the “theocentric paradigm of classical rationalism” whose core assumption is that cognition consists in conformity to a God’s-eye perspective (p. 115). 4 SW (I/7, p. 382). 5 Kant (1998, A84–85/B116–117); cf. AA (18, p. 267). 6 Ibid., A66–83/B91–115. 7 Ibid., A80–81/B106–107. 8 GA (IV/2, p. 6). 9 GA (IV/2, p. 8). 10 See Fichte, GA (IV/2, pp. 55, 113); SW (I, pp. 446, 449, 457–458, 462); SW (III, pp. 34, 35, 73); SW (IV, p. 49). 11 GA (IV/2, p. 6). 12 Kant (1998, B145–146). 13 Whether the absence of an absolute ground warrants criticism is one of the central debates of post-Kantian idealism. Fichte claims in the New Presentation that Kant’s metaphysical deduction at least implies a rigourous answer to the question quid facti: “the Critique of Pure Reason by no means lacks a foundation. Such a foundation is very plainly present; but nothing has been constructed upon it, and the construction materials—though already well prepared— are jumbled together in a most haphazard manner” SW (I, p. 479n). Contrast Henrich: “the quaestio juris can be answered in a satisfactory way even if the quaestio facti meets with insurmountable difficulties. Consider again the example of the last will: in many cases we are unable to produce a complete story of the way in which the will has been made. But if it can be determined in court that the will is authentic and valid, by means of only a few but crucial aspects, the question of right can still be answered decisively” Henrich (1989, p. 36). 14 GA (IV/2, p. 7). 15 SW (I, p. 442). 16 SW (I, pp. 425, 445). 17 SW (I, p. 442). Compare Fichte’s 1812 lectures on transcendental logic: “[Kant] was not so disinclined as he ought to have been [toward general logic. . .]” and “had not recognized that his own philosophy requires that general logic be destroyed to its very foundation.” SW (IX, pp. 111–112). 18 SW (I, pp. 445–446). 19 See Fichte, GA (IV/2, p. 179); SW (I, p. 87); SW (III, pp. 2, 9); SW (IV, pp. 14–15). 20 See Fichte, SW (I, pp. 271, 305); SW (I, pp. 458, 495); SW (II, pp. 445–446); SW (III, p. 77); SW (IV, pp. 14, 37); GA (II/4, p. 103); GA (IV/3, pp. 342, 480–481). For an illuminating account of the connection between a pragmatic history of the mind and a genetic deduction, see Breazeale (2013, pp. 70–95). On the relation between Fichte’s genetic deduction and Kant’s metaphysical and transcendental deductions, see Bruno (2018). 21 SW (I, p. 495). Compare an unpublished passage from 1793: “Sensibility, understanding, reason, the faculty of knowledge, the faculty of desire—can one

Determinacy, Indeterminacy, and Contingency  79 demonstrate [. . .] the necessity of all these? More specifically, can the whole of philosophy be constructed upon a single fact?” GA (II/3, p. 26). 22 In the New Presentation, Fichte addresses the question of how to transition from the I, which is absolute and thus determined by nothing outside it, to determinate conditions. Since “nothing determinate can be derived from what is indeterminate,” and since the I is “the ultimate ground of all explanation,” Fichte concludes that the I must be determined “by its own nature” SW (I, pp. 440– 441). From this, we can infer that deriving the conditions of the possibility of the I’s realization is the I’s own activity of self-determination, i.e., its own transition to greater determinacy. 23 GA (IV/2, p. 38). 24 GA (IV/2, p. 44). Compare Fichte’s claim that reciprocal interaction is “the category of categories” GA (IV/2, p. 212). On the Maimonian roots of Fichte’s principle of determinability, see Breazeale (2013, pp. 42–69). 25 GA (IV/2, pp. 2, 51). 26 On the proto-Hegelian character of Fichte’s derivational method in Foundations of the Entire Wissenschaftslehre, see Neuhouser (2014). 27 SW (III, p. 77). 28 Ibid., pp. 1–2. 29 Ibid., p. 2. While Fichte’s deduction supports Kant’s response to Humean skepticism, it also responds to Maimonian skepticism. Kant’s transcendental deduction shows that experience is impossible without the categories. While this proves our right to the categories, contra Hume, it raises the question of whether we actually apply them, i.e., whether they have reality, a challenge raised by Maimon (2010, p. 42). Fichte explicitly responds to this challenge when he says that, in deriving the categories from the I, the latter’s reality is “transferred” to the former; SW (I, p. 99); cf. SW (I, p. 121n); GA (IV/2, p. 8). Hence, he calls the categories “necessary actions” in Natural Right. This has the effect of both imbuing otherwise empirical phenomena like mutual address and bodily movement with transcendental significance and expanding ‘transcendental’ to denote certain actions. On the actuality problem raised by Maimon, see Franks (2005, pp. 243–249). 30 SW (III, p. 18). 31 Ibid., p. 19. 32 Ibid., p. 19. Gottlieb (2015) rightly argues that Fichte’s deduction of the external world is motivated by ethical rather than epistemic skepticism. 33 SW (III, pp. 24, 27, 29). 34 Compare Hegel: “the familiar, just because it is familiar, is not cognitively understood” (1977a, p. 18). 35 SW (III, pp. 32–33). 36 SW (III, p. 32). 37 SW (III, p. 33). 38 SW (III, p. 33). 39 See Franks (2016, p. 100). 40 SW (III, pp. 33, 36). 41 SW (III, p. 34). 42 SW (III, p. 42). 43 For an account of how, according to Fichte’s German idealist agenda, secondperson reference between rational subjects derives from a first principle, see Bruno (forthcoming a). 44 GW (19, §42R). 45 SW (I, p. 442). Compare Fichte’s criticism that Kant’s metaphysical deduction is inductive with Hegel’s criticism in the Science of Logic: “Kant made the profound observation that there are synthetic principles a priori, and he recognized

80  G. Anthony Bruno as their root the unity of self-consciousness, hence the self-identity of the concept. However, he takes the specific connection, the relational concepts, and the synthetic principles, from formal logic as given; the deduction of these should have been the exposition of the transition of that simple unity of self-consciousness into these determinations and distinctions; but Kant spared himself the effort of demonstrating this truly synthetic progression, that of the self-producing concept” GW (12, p. 205). 46 GW (19, §78). 47 GW (11, p. 388). 48 SW (I, p. 46). 49 SW (I, pp. 425–426). 50 Ibid., p. 422. For an account of why philosophy’s first task and first demand are, for Fichte, one and the same, see Bruno (forthcoming a). 51 GA (IV/2, p. 17). 52 GW (19, §78). 53 GW (4, pp. 389–390). 54 Ibid., p. 389. 55 Ibid., p. 393. Hegel chides Fichte for masquerading the I’s “infinite poverty” as an “infinite possibility of wealth.” Ibid. (p. 390). Compare his criticism of sensecertainty; W (3, p. 82). 56 Equiprimordially presupposing the I and the not-I is what necessitates the infinite practical striving for their unity in the Foundations, which inspires Hegel’s longstanding criticism that the Wissenschaftslehre guarantees its own irresolution (see, e.g., GW [4, pp. 45, 400–402]; [19, §60A2, §94A]; [21, pp. 123, 150, 227]), an incompleteness that Martin (2007) argues instantiates Hegel’s concept of bad infinity. 57 GW (21, p. 11). 58 See Fichte: “What then is the overall gist of the Wissenschaftslehre, summarized in a few words? It is this: Reason is absolutely self-sufficient; it exists only for itself” SW (I, p. 474). Compare Pippin: “If there is a ‘monism’ emerging in the post-Kantian philosophical world, the kind proposed by Fichte (and that decisively influenced Hegel [. . .]) is what might be called a normative monism, a claim for the ‘absolute’ or unconditioned status of the space of reasons” Pippin (2000, p. 164). 59 GW (19, §60A2). 60 SW (I, pp. 248, 279). 61 GW (19, §79). 62 Italics added. Compare Hegel: “what seems to happen outside of [Spirit], to be an activity directed against it, is really its own doing, and Substance shows itself to be essentially Subject” W (3, p. 39). 63 GW (19, §80–82). 64 For an account of why Fichte’s genetic deduction is an answer to both the question quid juris and the question quid facti, see Bruno (2018). 65 GW (21, pp. 68–69). 66 Ibid., p. 69. 67 Compare Hegel: “in speculative thinking [. . .] the negative belongs to the content itself” Hegel (1977a, p. 36). 68 GW (21, pp. 69–70). 69 Ibid., p. 236. According to Franks, “[Fichte’s] idea of intellectual intuition and [Hegel’s] idea of determinate negation are both attempts to conceptualize the same thing: the relationship between the ens realissimum or absolute first principle, and the fundamental forms or categories in virtue of which all possible entities may be determined and individuated” Franks (2005, p. 340).

Determinacy, Indeterminacy, and Contingency  81 70 For an account of how this question structures both Schelling’s philosophical development and his influential critique of Hegel, see Bruno (forthcoming b). 71 Compare Schelling: “I ask again, why is there a realm experience at all? Every reply I give to this already presupposes the existence of a world of experience. In order to be able to answer this question we should first of all have to have left the realm of experience; but if we had left that realm the very question would cease” SW (I/1, p. 310). 72 SW (I/1, pp. 162, 181). 73 SW (I/7, p. 360). 74 SW (I/6, p. 143). Further Presentations from the System of Philosophy expresses the union of Schelling’s earlier doctrines with the claim that intellectual intuition, philosophy’s “first cognition,” “snatches the ultimate doubling {of the real and ideal} away from the dualism it inhabits and establishes absolute idealism for the partial idealism of the world of appearances” SW (I/4, p. 404). 75 SW (I/7, p. 353). 76 Ibid., p. 352. 77 See Fichte: “idealism begins with a single basic law of reason, which it immediately establishes within consciousness. In order to do this, it proceeds as follows: it summons the listener or the reader to think freely of a certain concept. If he indeed does this, he will discover that he is obliged to proceed in a certain way. Here we have to distinguish between two different things: [1] The requested act of thinking, which can only be performed freely. The person who does not perform this act on his own will not be able to see any of the things set forth in the Wissenschaftslehre. [2] The necessary manner in which this free act of thinking has to be performed if it is to be performed at all. The basis for this necessity lies in the very nature of the intellect itself and is not a matter of free choice. This is something necessary, even though it only occurs in and by means of a free action. It is something discovered, even though its discovery is conditioned by freedom” SW (I, p. 445). 78 SW (I/7, pp. 359–360). 79 For an account of this criticism, see Bruno (2016). 80 SW (I/7, p. 360). 81 SW (I/7, p. 382). 82 Ibid., p. 382. 83 Schelling’s conception of will affords an account of evil that steers between the Scylla and Charybdis of sensation and reason. In Religion within the Boundaries of Mere Reason, Kant observes that evil poses a daunting question, for its ground lies neither in the senses, for which we are not responsible, nor in reason, which cannot “extirpate” the dignity of the moral law. Kant (AA 6, p. 35). On Schelling’s view, evil’s imputability is only explained by the positive idea of will as “the capacity for good and evil” SW (I/7, p. 352). 84 Ibid., p. 407. Compare Schelling: “no one has chosen [one’s] character following reasoning or reflection. One did not consult oneself” SW (I/8, p. 304). 85 GW (21, p. 56). 86 Compare Nietzsche (2006): “there is no ‘presuppositionless’ science—the very idea is unthinkable, paralogical: a philosophy, a ‘faith’ must always be there first, so that from it science can acquire a direction, a sense, a limit, a method, a right to exist” (Third Essay, §24). 87 SW (I/7, p. 359). 88 SW (I/7, pp. 383, 385). 89 SW (I/7, pp. 385, 407). 90 SW (I/9, p. 93). Compare Schelling: “most people are frightened [. . . by] abyssal freedom in the same way that they are frightened by the necessity to be utterly one thing or another” SW (I/8, p. 304).

82  G. Anthony Bruno

Works Cited Allison, Henry. 2006. “Kant’s Transcendental Idealism,” in A Companion to Kant, edited by Graham Bird (pp. 111–124). Oxford: Blackwell. Breazeale, Daniel. 2013. Thinking Through the Wissenschaftslehre. Oxford: Oxford University Press. Bruno, G. Anthony. forthcoming a. “The Thought of a Principle: Rödl’s Fichteanism,” in The Bloomsbury Companion to Fichte, edited by Marina Bykova. New York, NY: Bloomsbury. ———. forthcoming b. “The Facticity of Time: Conceiving Schelling’s Idealism of Ages,” in Schelling’s Philosophy: Freedom, Nature, and Systematicity, edited by G. Anthony Bruno. Oxford: Oxford University Press. ———. 2016. “ ‘As from a State of Death’: Schelling’s Idealism as Mortalism.” Comparative and Continental Philosophy 8(3): 288–301. ———. 2018. “Genealogy and Jurisprudence in Fichte’s Genetic Deduction of the Categories.” History of Philosophy Quarterly 35(1): 77–96. Fichte, Johann Gottlieb. 1962–2012. Gesamtausgabe der Bayerischen Akademie der Wissenschaften (cited as GA), edited by Reinhard Lauth, Hans Jacobs, and Hans Gliwitzky. Stuttgart-Bad Canstatt: Frommann-holzboog. ———. 1965. Sämmtliche Werke (cited as SW), edited by Immanuel Hermann Fichte. Berlin: de Gruyter. ———. 1988. Early Philosophical Writings, translated by Daniel Breazeale. Ithaca, NY: Cornell University Press. ———. 1994. An Attempt at a New Presentation of the Wissenschaftslehre in Introductions to the Wissenschaftslehre and Other Writings, translated by Daniel Breazeale. Indianapolis, IN: Hackett. ———. 1998. Foundations of Transcendental Philosophy (Wissenschaftslehre) nova methodo, translated by Daniel Breazeale. Ithaca, NY: Cornell University Press. ———. 2000. Foundations of Natural Right, translated by Michael Bauer. Cambridge: Cambridge University Press. Franks, Paul. 2005. All or Nothing: Systematicity, Transcendental Arguments, and Skepticism in German Idealism. Cambridge, MA: Harvard University Press. ———. 2016. “Fichte’s Kabbalistic Realism: Summons as ẓimẓum,” in Fichte’s Foundations of Natural Right: A Critical Guide, edited by Gabriel Gottlieb pp. 92–116. Cambridge: Cambridge University Press. Gottlieb, Gabriel. 2015. “Fichte’s Deduction of the External World.” International Philosophical Quarterly 55(2): 217–234. Hegel, G. W. F. 1968. Gesammelte Werke (cited as GW), Deutsche Forschungsgemeinschaft. Hamburg: Meiner. ———. 1970. Werke in 20 Bänden (cited as W), edited by Eva Moldenhauer and Karl Markus Michel. Frankfurt: Suhrkamp. ———1977a. Phenomenology of Spirit, translated by Arnold V. Miller. Oxford: Oxford University Press. ———. 1977b. Faith and Knowledge, translated by H. S. Harris and W. Cerf. Albany, NY: SUNY Press. ———. 1977c. The Difference Between Fichte’s and Schelling’s System of Philosophy, translated by H. S. Harris and W. Cerf. Albany, NY: SUNY Press. ———. 1991. The Encyclopedia Logic (with the Zusätze), translated by T. F. Geraets, W. A. Suchting, and H. S. Harris. Indianapolis, IN: Hackett.

Determinacy, Indeterminacy, and Contingency  83 ———. 2010. Science of Logic, translated by George di Giovanni. Cambridge: Cambridge University Press. Henrich, Dieter. 1989. “Kant’s Notion of a Deduction and the Methodological Background of the First Critique,” in Kant’s Transcendental Deductions: The Three ‘Critiques’ and the ‘Opus postumum, edited by Eckart Förster (pp. 47–68). Stanford: Stanford University Press. Kant, Immanuel. 1900. Kants gesammelte Schriften (cited as AA, except for the Critique of Pure Reason, cited in the A/B pagination for the 1781/1787 editions). Berlin: de Gruyter. ———. 1996. Religion and Rational Theology, translated and edited by Allen W. Wood and George di Giovanni. Cambridge: Cambridge University Press. ———. 1998. Critique of Pure Reason, translated and edited by Paul Guyer and Allen W. Wood. Cambridge: Cambridge University Press. Maimon, Salomon. 2010. Essay on Transcendental Philosophy, translated by Nick Midgley, Henry Somers-Hall, Alistair Welchman, and Merten Reglitz. London: Continuum. Martin, Wayne. 2007. “In Defense of Bad Infinity: A Fichtean Response to Hegel’s Differenzschrift.” Bulletin of the Hegel Society of Great Britain 28(55): 168–187. Neuhouser, Frederick. 2014. “Fichte’s Methodology in the Wissenschaftslehre (1794–95),” in The Palgrave Handbook of German Idealism, edited by Matthew C. Altman (pp. 400–419). Basingstoke: Palgrave Macmillan. Nietzsche, Friedrich. 2006. On the Genealogy of Morality, translated by Carol Diethe. Cambridge: Cambridge University Press. Pippin, Robert B. 2000. “Fichte’s Alleged Subjective, Psychological, One-Sided Idealism,” in The Reception of Kant’s Critical Philosophy: Fichte, Schelling, and Hegel, edited by Sally Sedgwick (pp. 147–170). Cambridge: Cambridge University Press. Schelling, Friedrich Wilhelm Joseph. 1856–1861. Friedrich Wilhelm Joseph Schellings sämmtliche Werke (cited as SW), edited by K. F. A. Schelling. Stuttgart: Cotta. ———. 1980. The Unconditioned in Human Knowledge: Four Early Essays 1794– 1796, translated by Fritz Marti. Lewisburg: Bucknell University Press. ———. 2000. Ages of the World, translated by Jason M. Wirth. Albany, NY: SUNY Press. ———. 2001. “Further Presentations from the System of Philosophy, translated by Michael Vater.” Philosophical Forum 32(4): 373–397. ———. 2006. Philosophical Investigations into the Essence of Human Freedom, translated by Jeff Love and Johannes Schmidt. Albany, NY: SUNY Press.

3 Free Thinking in Schelling’s Erlangen Lectures Gregory S. Moss

Introduction In Schelling’s Erlangen lecture Über die Natur der Philosophie als Wissenschaft, Schelling makes the astounding claim that philosophy is not a demonstrative science. Schelling claims that “philosophie ist nicht demonstrative Wissenschaft.” Instead, philosophy is a “Freie Geistesthat.” Philosophie is “das Freie Denken.” Schelling could not be more unequivocal when he baldly claims that philosophy is “ausdrücklich ein nicht Wissen”; indeed, it is “ein Aufgeben alles Wissens.” Yet, in this short lecture on the nature of philosophy as science, Über die Natur der Philosophie als Wissenschaft, Schelling’s intention is to lay bare the nature or the essence of philosophy as a science, as a Wissenschaft. Since philosophy is itself not a science, but free thinking, insofar as philosophy is undertaken and conceived as a science, it is false and in a state of self-alienation, for its determination does not correspond to what it is in itself. As is well known, “crisis” stems from the Greek word for “separation.” In Wissenschaft, philosophy is in crisis, for it is separated from itself. Schelling’s Erlangen lecture is, therefore, an inquiry into the crisis that is constitutive of philosophy as science. In the following, I show that indeterminacy plays a fundamental role in Schelling’s account of the crisis of philosophical knowing and the character of free thinking. In the Erlangen lectures, rather than propose a positive or negative conception of indeterminacy, each of which is a relative conception, Schelling argues that free thinking consists in an act of intellectual ecstasy by which philosophy transcends all concepts, and Absolute Indeterminacy is revealed. By construing philosophy as free thought, Schelling contraposes free thinking to philosophy as science. In the course of the chapter, I will show that, on Schelling’s own terms, opposing free thinking to science entails that free thought fails to be free, and becomes a kind of Wissen. For this reason, Schelling’s propositional construal of philosophy as free thought is self-­ contradictory. Rather than shy away from this self-contradiction, I argue that this self-contradiction is necessary in order to defend Schelling’s insight that philosophy is not a science, but free thinking.

Free Thinking in Schelling’s Erlangen Lectures  85

Philosophy and Science The task of metaphysics is to provide a system of Being qua Being. Since metaphysics is the science of first principles, all other forms of knowing depend upon metaphysics for their being. Since metaphysics exists, or is actual at least as a practice, it is also possible. Given that metaphysics is possible as a practice, we can ask, “What makes metaphysics possible?” In addition, we can also ask whether metaphysics is possible as a science, or Wissenschaft. Without inquiring into the possibility of metaphysics, metaphysics runs the risk of falling victim to dogmatism. Since all other forms of knowing depend upon metaphysics, no other science or form of knowing can demonstrate how metaphysics is possible. Since it is metaphysics that constitutes the inquiry into first principles, it must fall to metaphysics to inquire into how metaphysics is possible. Hence, metaphysics is not only an inquiry into Being qua Being, but it is also an inquiry into its own knowing of Being qua Being. Insofar as this inquiry is characterized as knowledge, it can be construed as the knowledge of knowledge of Being. For this reason, the inquiry into first principles can equally be construed as epistemology, since it is an inquiry into the very possibility of the knowledge of first principles. Meta-metaphysics is the discipline that constitutes the knowledge of metaphysics itself, and thereby takes metaphysics as its object. We should note at the outset that meta-metaphysics is not a discipline distinct from metaphysics, but belongs to the activity and content of metaphysics proper. By inquiring into whether metaphysics is possible as a science, as well as the inquiry into the condition for the possibility of metaphysics as a practice, metaphysical inquiry takes itself as its own object. Naturally, this makes metaphysics an inherently reflexive discipline. As a reflexive discipline inquiring into the possibility of metaphysics as a science, meta-metaphysical inquiry aims to uncover how the relationship between knowing and Being ought to be construed. For this reason, meta-metaphysical inquiry is not only interested in the discipline of metaphysics, namely the knowing of Being, but by virtue of its inquiry into the knowing of Being it aims to establish some truth about Being itself, namely whether Being, as the object of metaphysics, is itself structured in such a way as to be an object of philosophical knowing. Accordingly, meta-metaphysics also aims to establish whether Being as it is in itself is amenable to the structure of knowing. Therefore, within the meta-metaphysical aspect of metaphysics alone one discovers an identity of subject and object, for meta-metaphysics itself contains the two-sides of metaphysical inquiry: an inquiry into both the activity of the subject, or (i) the knowing of Being, as well as (ii) the object of that inquiry, Being itself. Most generally, we have established two theses: metaphysics is meta-metaphysics, and meta-metaphysics is metaphysics. Schelling’s Über die Natur der Philosophie als Wissenschaft is an investigation into the very concept of a system.1 By determining what it is for philosophy to be a Wissenschaft, Schelling is aiming to provide a meta-theory,2 a metametaphysics. Schelling asks about the principle of the possibility of the system.3

86  Gregory S. Moss Given that meta-metaphysics is constituted by both a metaphysical and epistemological side, we can characterize the subject matter of the inquiry in either or both ways. The subject of philosophical meta-theory or meta-metaphysics is universal. Indeed, the inquiry is into that which “durch alles geht.”4 On the ontological side, Being is indeed held in common by all beings, and “goes through them all.” Likewise, the knowledge of our knowledge of being is also universal, for it concerns the structure of knowledge as such. Accordingly, meta-metaphysics is an inquiry into what is universal. But it is also unconditioned and “muß durch alles gehen und in nichts bleiben.”5 Insofar as Being is universal, it includes all determinations and cannot thereby be conditioned or limited by any one of them or set of them. Likewise, since meta-metaphysics is a knowing of all knowledge of being, it is not itself conditioned by any other knowledge. Accordingly, both in respect to the epistemic and ontological sides, the subject matter of meta-metaphysics cannot be limited by any of its determinations, for it runs through and surpasses them all. Accordingly, the subject of the philosophical meta-theory is unconditionally universal or what is the same: Absolute Being and Absolute Knowing. Since each of these is not limited by any determinations, Absolute Being and Absolute Knowing are infinite. Schelling is clear that there is only one subject of all determinations, not many.6 For Schelling, rather than conceive of knowing and Being as two Absolute subjects, they express two different sides of the same Absolute unity.7 This Absolute subject has been traditionally understood to be the principle of philosophy.8 The insight that the Absolute is not finite can be grasped whether the Absolute is conceived in epistemological or ontological terms. Being as such includes all beings. Accordingly, there is no being external to Being by which Being could in principle be constrained. Hence, Being is not finite; it is infinite. For this reason, when reflecting on the Absolute, one ought not think of it as a being, or as “ein seyendes,” for all beings are differentiations of Being, which must necessarily exceed them. Schelling is clear that the Absolute is not finite: Hier muss alles Endliche, alles, was noch ein Seyendes ist, verlassen werden.9 In sum, philosophy is concerned not only with knowledge of Absolute Being but also knowledge of Absolute Knowing, and these appear to be infinite. Philosophy is concerned with the knowing of knowing and must thereby establish for itself what philosophy is and must distinguish itself from other disciplines.10 Schelling is clear that philosophy’s subject matter cannot be contained in any form: Es ist in einer unaufhaltsamen Bewegung, in keine Gestalt einzuschliessen, das Incoercible, das Unfassliche, das wahrhafte Unendliche.11

Free Thinking in Schelling’s Erlangen Lectures  87 In addition to being absolutely infinite, the object of philosophical knowledge is also incomprehensible and incoercible. Indeed, to define and conceptualize means to set something within definite limits: Was heiβt definieren? In bestimmten Grenzen einzuschliessen.12 When one claims that “S is P,” one confines S to P and implicitly denies that insofar as it is P that it could be not-P. For this reason, to conceive by means of concepts is to limit the subject matter and to transform what is conceived into the finite. “S” is determinate insofar as it has its own independent content “P,” which excludes its negation “not-P.” Determinacy entails finitude and relativity. By conceiving of the subject as finite, one cannot conceive of Absolute Knowing or Absolute Being, for these are infinite. Because only that which can be grasped by means of concepts is conceivable, it appears that Schelling identifies the true infinite with the incomprehensible and indefinable: Ich muss eben das indefinable, das nicht zu definirende des Subjekts selbst zur Definition machen.13 Further, since conceptual determinations cannot in principle encompass the Absolute, they are only ever able to connote what is relative. By means of concepts, it appears that one can only know relative knowing, and relative being. Philosophy, insofar as it is concerned with the Absolute, is not primarily concerned with a determinate “something,” for to be a determinate something is to be finite, and to be opposed to an other. Yet, philosophy is not concerned with a mere nothing, either: its primary concern is ontological, rather than ontic. In terms of beings, concepts only grasp differentiations or determinations of being, but not Being itself. In respect to knowing, concepts are only capable of grasping particular differentiations or predicates, but not the subject of all predication or the genus of all predicative determinations. Given that philosophical thinking attempts to know Being by means of concepts, it follows that metaphysical and epistemic thinking about the absolute necessarily leads to aporia. Skepticism appears to be the final result of metaphysical thinking, and one of the fundamental impasses that lies as a stumbling block to philosophical knowing. Of course, the history of philosophy seems to indicate that no progress has been made in metaphysics: metaphysical knowledge does not progressively build on itself, as the sciences do. For this reason, the development of a system of philosophical knowledge of the Absolute appears to be futile. Indeed, Schelling appears to have gone further: skepticism is not contingently related to the pursuit of metaphysical knowledge, but necessary to and inherent within metaphysics itself.14 Meta-metaphysics recognizes this skepticism as constitutive of metaphysics itself.

88  Gregory S. Moss The impasse can be best illuminated by Schelling’s own claim that the Absolute is infinite. By imposing the concept of the infinite on the Absolute, and thereby defining the Absolute as infinite, the philosopher opposes the Absolute to finitude. Insofar as the Absolute is opposed to finitude, it is not truly infinite, for it is limited by the finitude which it excludes. The Absolute is determined to be infinite and is thereby conceived by means of concepts. Schelling claims that the concept of the “unendlich” only expresses a negative concept, a negation of Endlichkeit. By defining the Absolute as infinite and incomprehensible it becomes finite and comprehensible.15 The deep irony here is that human knowledge destroys what it wants by wanting it.16 The knowing of the Absolute thereby turns back on itself and the incomprehensible becomes comprehensible and the infinite becomes finite: Absolute Subject=indefinible, das Unfassliche, das Unendliche. [. . .] Es ist nicht so indefinible, dass es nicht auch ein Definibles werden könnte, es ist nicht so unendlich, dass es nicht auch endlich werden könnte, nicht so unfasslich, dass es nicht auch fasslich.17 Since the Absolute transcends both finitude and relativity, the Absolute transcends all determinacy, whether this be conceived in epistemic or ontological terms. Accordingly, the Absolute is indeterminate. But insofar as the Absolute is conceived as indeterminate, the Absolute is posited as standing in contrast and in relation to the determinate. Accordingly, the indeterminate character of the Absolute becomes determinate, and the Absolute becomes relative to an external factor. Thus, in virtue of defining the Absolute as indeterminate, the Absolute becomes determinate and relative. For Schelling, any attempt to think the Absolute falls into contradiction.18 Indeed, that the principle is neither B nor not-B is a violation of excluded middle, or insofar as it is just as much B as it is not-B, a violation of the principle of non-contradiction. Schelling is clear that this is the case “with every other determination.”19 Another term which Schelling employs to describe the Absolute in this essay is Gleichgültigkeit or indifference. The Absolute is equally applicable to both determinations without preference or priority. By determining it as either B or not-B, one renders the Absolute finite and relative, and subordinates it to the indifference that transcends either determination. The meta-metaphysical investigation has uncovered the inherent conflict between the form of the concept and the Absolute: Man muss sich überzeugt haben, dass dieser Widerstreit einen objecktiven Grund hat, dass er in der Natur der Sache selbst, in den ersten Wurzeln alles Daseyns gegründet ist.20 The self-undermining character of this dialectical procedure brings the Absolute to a contradiction. Since it would appear that attempting to systematize

Free Thinking in Schelling’s Erlangen Lectures  89 the Absolute undermines itself, by thinking through the concept of a system one transcends each particular system. 21 As Gabriel notes, rather than succumb to skepticism and abandon all attempts to develop a system of philosophy, Schelling incorporates the skeptical posture toward metaphysics into the very concept of metaphysics itself.22 The conflict between metaphysical positions makes possible the idea of the system: Also die Idee des Systems überhaupt setzt den notwendigen und unauflöslichen Widerstreit der Systeme voraus: ohne diesen würde sie gar nicht entstehen.23 No one metaphysical system can dominate over the others exactly because conceptual structure only applies at the level of relative differentiations, and is not endemic to the Absolute as such.24 From these reflections, we can summarize the five conditions that Schelling lays out for a system of philosophy. First, it must recognize a fundamental conflict [Streit] in knowing. Second, this conflict must be made manifest [both historically and systematically] and be pursued in all directions. Third, the conflict is grounded in first principles. Fourth, one must give up hope to master the whole with one exclusive system. Fifth, one must not attempt to unify all systems into one overarching unity.25 Although Schelling’s meta-metaphysical inquiry has shown that skepticism lies waiting in the heart of metaphysics, Schelling also shows how integrating it into the very practice of metaphysics overcomes the skeptic’s challenge. By attempting to think the Absolute in conceptual terms, the indeterminate becomes determinate, and the infinite finite. As a result of the failure to capture the Absolute in conceptual form, the Absolute is revealed as that which is (i) immanent in all determinations and (ii) transcends all determinations. As Schelling writes: Denn nur indem es Gestalt annimmt, aber aus jder wieder siegreich heraustritt, zeigt es sich als das an sich Unfassliche, Unendliche.26 As Gabriel translates this passage, “only by assuming a shape, yet at the same time victoriously transcending that shape does it show itself to be the ungraspable, the infinite.”27 As long as the infinite is limited to the concept of the infinite, the indeterminate to the concept of the indeterminate, or the incomprehensible is constrained by the concept of the incomprehensible, the infinite is not itself infinite, and cannot be known as infinite or incomprehensible. But insofar as the infinite breaks out of the conceptual determination imposed upon it, namely in and through the failure of the concept to constrain the infinite, the infinite reveals itself to be what it is: that which transcends all conceptual determination and thereby all finitude. Only insofar as the

90  Gregory S. Moss indeterminate breaks out of conceptual determination in and through the failure of the concept to determine the Absolute, can the Absolute reveal itself to be indeterminate. Through the failure of capturing the Absolute with the concept, the Absolute reveals itself to be truly infinite and indeterminate. Indeed, for this reason it is exactly the skeptical posture by which the infinite can reveal its true self. Through the skeptical posture that demonstrates the impossibility of knowing the Absolute by means of concepts, one indirectly reveals the indeterminacy of the Absolute: that which encompasses all conceptual and finite determinations while simultaneously superseding them. Indeed, by appropriating the Absolute as an object of conceptual thinking, the subject makes the Absolute into an object. This process of transforming the Absolute into an object for conceptual or discursive knowing renders the Absolute a contradiction. Schelling asks: Wie können wir jenes absolute Subjekt, die ewige Freiheit wissen; dieser Frage liegt die noch allgemeinere zu Grunde: Wie kann sie überhaupt gewußt werden? Nämlich: 1) Es ist ein Widerspruch darin, daß die ewige Freiheit erkannt werden soll. Sie ist absolutes Subjekt = Urstand; wie kann sie denn Gegenstand werden? Unmöglich kann sie es werden als absolutes Subjekt, denn als solches steht sie zu nichts in gegenständlichem Verhältniß.28 The Absolute cannot appear as it is “an sich” when it is taken as an object of knowledge.29 But one must have access to what the infinite is in order to infer that one’s judgment about the infinite is false. As is clear from the passage, the Absolute Subject is here identified with Absolute Freedom. By revealing the Absolute Subject to be that which transcends form, one reveals the Absolute Subject to be that which is free from all determination. What is more, the Absolute is free because it is not only free from all determination, but because it is free to be enclosed within or immanent in every determination: Aber ursprünglich ist es doch frei, sich in eine Gestalt einzuschließen und nicht einzuschließen.30 The revelation of the Absolute as the true infinite is the revelation of the Absolute as Eternal and Absolute Freedom. Indeed, Schelling baldly identifies Eternal Freedom with the Essence of the Subject.31 The Absolute as Absolute cannot be determined by any particular form, but is that which is capable of taking on a particular form. Or what is the same: every form and differentiation is a differentiation and limitation of Absolute freedom. In this way, there is nothing antecedent to the Absolute which could determine it. The freedom of the Absolute is its Absolute indeterminacy: it consists in the fact that it has nothing to rely upon.32

Free Thinking in Schelling’s Erlangen Lectures  91 To put it tersely, Gabriel writes that for Schelling metaphysical knowledge is only actual in its self-destruction:33 Nun sehen wir freilich, dass sie in nichts bleibt, jede Form wieder zerstört34 Through the self-destruction of every determination and system of the Absolute, the Absolute is revealed as it is an sich: absolutely indeterminate. Schelling characterizes the Absolute as Die Ewige Magie. It is that which is, yet transcends conceptualization. In this sense, the Absolute is exactly that which cannot be explained. It cannot be denied, for it is a condition for the very existence of everything that is conceptually intelligible and determinable, but it is itself beyond explanation. Indeed, the crisis of human knowledge may be described as the conflict that ensues between human knowing and Absolute freedom. As Schelling writes: Also indem der Mensch jene ursprüngliche Freiheit sich zum Objekt macht, es mit ihr zum Wissen bringen will, entsteht nothwendig folgender Widerspruch: er will die ewige Freiheit als Freiheit wissen und empfinden, aber indem er sie zum Gegenstand macht, wird sie ihm unter der Hand zur Nichtfreiheit, und doch sucht und will er sie als Freiheit.35 “Wissen” or knowledge appears here to be characterized by demonstrative knowing. Proofs, demonstrations, and arguments more generally are constituted by judgments. Judgments are constituted by concepts, which if we follow Kant, are the logical functions by which particulars are unified. By taking Absolute Freedom as an object of Wissen or knowledge, Absolute freedom ceases to be free. The Absolute can only appear as an object of knowledge as a contradiction. Accordingly, the very search for Absolute Freedom renders it inaccessible to the knower. For this reason, the knower has alienated himself from Absolute Freedom by attempting to know it. Absolute Freedom is beyond knowledge and can only appear in Wissen as a kind of self-alienation. On the one side, there is determinacy, both the determinacy of knowing and its ontological counterpart. One the other side, there is the indeterminate Absolute, or the absolutely Indeterminate, which eschews all determination. Freedom is alienated from itself in Wissen. Since the freedom of the Absolute has both ontological and epistemic sides, not only is the freedom of Absolute Knowledge alienated from itself, but Absolute Being is also in a state of self-alienation. This condition is a “zerreißendsten Zweifel” and “ewigen Unruhe.” Insofar as philosophy is the perpetual search for this absolute knowing, which perennially escapes it, philosophy takes on a Socratic form.36 The crisis of knowing and being is the separation of the Absolute from itself: As long as the Absolute is constrained by the form of knowledge,

92  Gregory S. Moss the human being does not know the Absolute, and does not have absolute knowledge. For this reason, Wissen is a crisis of separation between human knowing, conceptual and relative knowing, and Absolute knowing. It is the principle of separation by which Absolute knowing appears inaccessible to human thinking. Wissen is the principle by which Absolute Being remains inaccessible to human knowledge, in which the Absolute only ever appears as some relative differentiation, some being among beings. On the one side of the separation there is human consciousness, which is constituted by an absolute ignorance, or Absolute Non-knowing, “absoluten Nichtwissens” about the Absolute. On the other side of the crisis there is the Absolute Subject which is that which is recognized as that which is not known by human consciousness. It is this crisis in which knowing is separated from the Absolute which is the source of error.37 What is the status of philosophy in respect to the Absolute? Philosophy is not Wissen: Philosophie ist nicht demonstrative Wissenschaft.38 Indeed, philosophy is “ausdrücklich ein nicht Wissen.” It is the act of giving up on Wissen. Philosophy is “ein Aufgeben alles Wissens für den Menschen.” The Absolute only appears when the subject ceases to take the Absolute as an object: So lang er noch wissen will, wird ihn jenes absolute Subjekt zum Objekt werden, und wird es eben darum nicht an sich erkennen. Jenes absolute Subjekt is nur da, sofern es nicht zum Gegenstande macht.39 Instead of scientific knowing, philosophy ought to be construed as “Freie Geistesthat” and “das Freie Denken.”40 Philosophical thinking begins exactly where knowing leaves off: Denken ist aufgeben von Wissen.41 Given that all knowledge, i.e., Wissen, is contingent upon conceptual form and structure, it follows that thinking the Absolute cannot be knowledge; it cannot be Wissen. Knowing appears to have its application and domain in the relative and finite determinations and beings that are amenable to conceptual differentiation and predication. Philosophy is the thinking that happens when the crisis separating human knowing from the Absolute is confronted.42

Free Thinking, Mysticism, and Indeterminacy Free thinking is that thinking by which Die Ewige Freiheit is revealed. In the attempt to apply concepts beyond the relative, the concept fails to capture

Free Thinking in Schelling’s Erlangen Lectures  93 what the Absolute is. Instead, the concept falls into its negation, and the philosopher experiences the tragedy of contradiction. By my reading, it is in the experience of the contradiction that free thinking consists.43 In the experience of contradiction, the philosopher experiences the domain where knowing no longer applies and non-knowing, nicht-Wissen reigns. Yet, the philosopher cannot simply turn away from the Absolute or deny that it is, for without it there would be no knowing at all. Accordingly, in the experience of contradiction, the philosopher is compelled to recognize the Absolute as the field of non-knowing that encompasses all knowledge. Indeed, to think freely means to dwell in or make a home in non-knowing. In virtue of concepts falling into their opposites, such as the conceivable falling into the inconceivable, the determinate into the indeterminate, the infinite into the finite, etc., the philosopher experiences that which transcends all conceptual and ontological boundaries and differentiations, and thereby experiences the Absolute. Of course, this experience cannot itself be characterized as conceptual. Rather, it is itself contradictory; for this reason it is an experience that itself cannot be explained in conceptual terms. Hence, it is an experience that can only be characterized as non-conceptual or as an experience of nonknowing, nicht-Wissen. Thus, through the self-destruction of metaphysical knowledge, Absolute Freedom is revealed in the experience of non-knowing, which is constitutive of free thought. Accordingly, free thinking is the thinking of Absolute and Eternal Freedom. Although reason fails to know the Absolute conceptually, the thinking by which the Absolute is revealed is impossible without an act of reason. This thinking does not consist in the successful correspondence of a concept with an object, for this would be Wissen; rather it consists in the non-conceptual disclosure of Absolute Being as well as one’s own Absolute non-knowing. By working through and beyond knowing one learns to think freely: In der Philosophie ist nicht der Mensch der Wissende, sondern er ist das dem eigentlich Wissenerzeugenden widerstrebende, durch beständigen Widerspruch es anhaltende—reflektirende—, aber eben darum für sich gewinnende freie Denken.44 Because philosophy is free thinking, insofar as the philosophical enterprise continues to be construed as knowing, philosophy is alienated from itself. As knowing, philosophy does not correspond to what it is in itself, and is in a problematic self-relation, which it must strive to overcome in the nonknowing that constitutes free thinking.45 It is only because insight into the completeness of the Absolute is achieved in das freie Denken that we can be confident in our assertion that philosophical Wissen is incomplete, and a moment of self-alienation. To put it simply, the non-conceptual experience of the Absolute in non-knowing reveals the Absolute in its completeness, and is the correlate to the incomplete knowing

94  Gregory S. Moss of the Absolute on the side of Wissen. Indeed, free thinking overcomes the perpetual Krisis from which Wissen cannot escape. In free thinking, one sees the whole of knowing, because one no longer perceives it from within the genus of the concept as such. Instead, by transcending each particular concept, one sees the whole genus as such, that is, one immediately grasps the concept as such.46 Although Gabriel points to the way that Schelling integrates skepticism into his conception of metaphysics in order to overcome the skeptical challenge, he does not mention the overtly mystical dimensions of Schelling’s characterization of philosophy as free thought. Given that all conceptual and therefore all scientific conceptions of the Absolute constrain philosophy and human knowing to relative knowing and relative being, in order to think the Absolute one must give up on knowing altogether: Wer wahrhaft philosophieren will, muß aller Hoffnung, alles Verlangens, aller Sehnsucht los seyn, er muß nichts wollen, nichts wissen, sich ganz bloß und arm fühlen, alles dahingehen, um alles zu gewinnen.47 To truly philosophize, or to think freely, one must give up on one’s desire to know. For as long as one desires knowledge of the Absolute, then one will not be able to think Absolute Freedom an sich. Of course one must desire to know the Absolute in order for one to err, and one must err in order for one to see that the Absolute cannot be known by means of concepts. But upon recognizing that the Absolute transcends concepts, one must give up on the desire to discover a system of conceptual determinations that truly corresponds to it. To think the Absolute by means of a concept, indeed to think by means of concepts generally is to relate to what is thought in a mediated way. Following Kant, it seems that for Schelling concepts are mediated modes by which we relate to what is thought. In the mediated relation to what is thought, the subject is not identifiable with what is thought. Instead, since the act of conception posits what is thought as something other to the subject, in virtue of the act of conceptualizing, the subject is distinct from what is thought, and relates to what is thought through a third term, such as the concept. Accordingly, to think what is thought by means of concepts is to take up what is thought as an object that is distinct from the subject. Throughout Schelling’s lecture On the Nature of Philosophy as Science, Schelling is clear that in Wissen the Absolute is taken as an object of knowing for a subject from which it is distinct. If to philosophize truly, or engage in free thought, one must transcend conceptual determinations, this appears to also entail that in free thinking one must transcend the mediated relation to an object. Hence, free thinking is not a mediated relation to a distinct Absolute object, but an immediate relation to the Absolute. According to Schelling, Das Ziel, also, ist das unmittelbare Wissen der ewigen Freiheit.48

Free Thinking in Schelling’s Erlangen Lectures  95 To think freely, or truly philosophize means to overcome mediated knowing and think the Absolute in an unmediated way. Accordingly, in order to cease to know, one must cease to take the Absolute as the object of knowing. By ceasing to relate to the Absolute as an object, one ceases to impose judgments or concepts on the Absolute. Accordingly, one abandons all conceptual differentiations and thereby also ceases to attend to the ontological differentiations that correspond to those conceptual differentiations or predicates. To truly philosophize means to abandon oneself to the infinite, indeterminate, and undifferentiated: Nur derjenige ist auf den Grund seiner selbst gekommen, und hat die Tiefe des Lebens erkannt, der einmal alles verlassen hatte, und selbst von allem verlassen war, dem alles versank, und der mit dem Unendlichen sich allein gesehen.49 In order to recognize the “depth of life” and to arrive at the “ground of the self,” one must have abandoned everything. Only in having abandoned everything can the infinite be recognized. By turning away from all predicative and ontological differentiations, one no longer relates to the Absolute as a being or as a concept, but to the Übergottheit: Es ist also insofern über Gott, und wenn selbst einer der vorzüglichsten Mystiker früher gewagt hat von der Übergottheit zu reden, so wird dies auch uns verstattet sein.50 Rather than conceive of the Absolute as a god to whom the world stands in contrast, the Absolute may be conceived as the Übergottheit that transcends that opposition. Only by becoming purified of all mediation and differentiation can what is unmediated and undifferentiated be revealed to the human subject. What is it to think freely? Thinking freely means allowing the mode of Wissen to destroy itself, thereby revealing the Absolute. Through the selfdestruction of conceptual knowing, the subject experiences the Absolute directly by ceasing to relate to the Absolute as an object of knowing. The direct non-conceptual experience of the Absolute is an immediate relation to the Absolute. Because this immediate experience transcends the principle in virtue of which differences are conceptualized, free thinking also entails becoming cleansed of all determinations: Denn nur dem Reinen offenbart sich das Reine. 51 Because free thinking ceases to relate to the object as something distinct from the subject, philosophy can only recognize the Absolute if the subject itself is the object, and the object is the subject.52 Just as much as the object cannot be distinct from the subject, so the subject cannot be posited

96  Gregory S. Moss as something distinct from the object. Otherwise, free thinking would be limited and unfree; it would fall back into the alienated form of knowledge. Accordingly, the immediate insight into the Absolute requires negating the self as a subject that is separate from what it thinks. To abandon oneself to the undifferentiated means negating the self. One cannot give up the objective relation to the Absolute without negating the independence of the subject as well. Schelling calls this Ekstase: Man hat dieses ganz eigenthümliche Verhältniß sonst wohl auszudrücken gesucht durch das Wort intellektuelle Anschauung. [. . .] Allein eben weil dieser Ausdruck erst der Erklärung bedarf, so ist es besser, ihn ganz bei Seite zu setzen. Eher könnte man für jenes Verhältniß die Bezeichnung Ekstase gebrauchen. Nämlich unser Ich wird außer sich, d. h. außer seiner Stelle, gesetzt. Seine Stelle ist die, Subjekt zu seyn. Nun kann es aber gegen das absolute Subjekt nicht Subjekt seyn, denn dieses kann sich nicht als Objekt verhalten. Also es muß den Ort verlassen, es muß außer sich gesetzt werden, als ein gar nicht mehr Daseyendes. Nur in dieser Selbstaufgegebenheit kann ihm das absolute Subjekt aufgehen in der Selbstaufgegebenheit, wie wir sie auch in dem Erstaunen erblicken.53 In this fascinating passage Schelling makes it clear that in the philosophical thinking of the Absolute, one relates to the Absolute immediately. Since this immediate relation is an act of thinking, and not a sensuous intuition, free thinking might be characterized as intellectual intuition. Schelling prefers to employ the term Ekstase. In the ecstatic relation of the self to the self, the I is set outside itself, außer sich. As an I that is opposed to and distinct from what it thinks, it is opposed to the Absolute subject, and is a separate being. In the ecstatic moment of free-thinking, the I ceases to be a separate and distinct subject that opposes the Absolute. For this reason, in ecstasy the separate I is outside itself or other to itself as a separate I. In virtue of giving up on the self, or Selbstaufgegebenheit, the self or I is merged into and with the Absolute subject, for there is no longer any mediating factor by which the Absolute could be held apart from the subject. In the experience of the unity of subject and object, freedom experiences its own groundlessness. The freedom of the Absolute now merges with and becomes indistinguishable from the agency of the subject. In free thinking, when one transcends concepts by means of concepts, the conceptual determination of the Absolute is set outside itself and knowing itself is transformed into not-knowing: in ecstasy knowing goes beyond itself into non-knowing. By predicating of the Absolute that it is “indeterminate,” one makes the Absolute determinate. Accordingly, the predicate “indeterminate” negates itself and fails to correspond to the Absolute. Yet, in that very failure of the predicate “indeterminate” to correspond to the Absolute, the Absolute is revealed to be indeterminate, for it reveals itself

Free Thinking in Schelling’s Erlangen Lectures  97 as eschewing all determinations, even the determination “indetermination.” Hence, free thinking is ecstatic—it reveals the indeterminate by means of its own self-transcendence. This ecstasy is free because it removes any and all determinations that might constrain its thinking of Absolute Freedom. Since the Absolute transcends all determinations, it is the indeterminate Absolute. Insofar as the absolutely Indeterminate does transcend determinacy, it is constituted by the absence of determinacy. The absolutely Indeterminate is negative: it stands against its opposite as an absence of determinacy. But it is also positive, for it is the being of the Absolute as such: it is that which is without qualification, yet cannot be reduced to any determination and that which makes any determination possible. In this way, its positivity consists in the fact that it stands alone as the one and only Absolute, which is affirmed in the experience of the contradiction. The absolutely Indeterminate does not stand in relation to any determination. For this reason, it would be equally problematic to conceptualize the absolutely Indeterminate as either positive or negative. Indeed, although each conception of the indeterminate has its own content, and in that respect is positive, each is also determined by its exclusion of the other, and is thereby negative. Indeed, the fact that each excludes the other (the one positive, the other negative) makes both finite determinations. Since each is a relative finitude, and the absolute Indeterminate transcends finitude and relativity, the absolutely Indeterminate cannot be identified with any conception of the positive or negative. What is more, by conceiving the Indeterminate to be absolute, one sets the Absolute in relation to what is relative, and thereby undermines its indeterminacy and its claim to be absolute. The absolutely indeterminate cannot be absolutely indeterminate. The absolutely indeterminate is contradictory: it is both positive and negative, and neither positive nor negative. In the contradiction the indeterminate Absolute shows itself to be absolutely Indeterminate only insofar as it is immanent in all determinations, e.g., both negatively and positively indeterminate, and transcends all determinations, e.g., neither positively nor negatively indeterminate. Indeed, the absolutely Indeterminate does not exclude all determinations; rather it includes them as determinations that have been overcome: it is a true, not a bad infinite in Hegel’s terminology. For Schelling in Erlangen, the proper object of philosophy is the Absolute and the Absolute is the absolutely non-conceptual Indeterminate that reveals itself in the ecstatic self-overcoming of free thought. To put it ecstatically: the positive significance of indeterminacy is revealed only in and through the failure of philosophy to conceptually determine the indeterminate to have positive significance. The positive significance of indeterminacy shows itself in the failure to express the meaning of its positivity. Not only does this ecstatic thinking remove all obstacles that stand between itself and the Absolute, but the form of its act is free. Because it negates itself, and consists in the self-removal of all concepts, ecstatic thinking is also self-determining and thereby free. Since ecstatic free thinking

98  Gregory S. Moss removes all mediating concepts, the Absolute is present immediately or intuitively in an intellectual act. Because it is ecstatic, it is not wrong to say that free thinking intuits the Absolute intellectually. Intellectual intuition posits the object by thinking it. Ironically, free thinking posits the Absolute in its failure to conceive it in non-knowing: In jener Selbstaufgegebenheit, jener Ekstasis, da ich als ich, mich erkenne als völliges Nichtwissen, wird mir unmittelbar jenes absolute Subjekt zur höchsten Realitaet. Ich setzte das absolute Subjekt durch mein Nichtwissen (in jener Ekstasis).54 Schelling’s concept of intellectual intuition in his Erlangen Lectures diverges significantly from his earlier conception in the System of Transcendental Idealism (1800). Despite various similarities, in the System intellectual intuition is not identified with ecstasy. Rather, in the System the self does not posit itself through ecstasy, the self-destruction of science, and absolute self-abandonment. In Erlangen, intellectual intuition is constituted by the complete abandonment of self and the self-destruction of scientific thinking. Nonetheless, the Absolute Subject does in fact appear in the Erlangen Lectures as it does in the System, for knowledge of the Absolute subject, and oneself as Absolute subject, the God beyond God, is realized by means of ecstasy. This is evident in the continuity of Schelling’s description of the Absolute as Indifference. In Erlangen intellectual intuition becomes thoroughly ironic: in the ecstatic abandoning of self and concept, the self and the concept are known. Given the identity of subject and object, the knowing of the object is knowledge of the subject. Accordingly, the self-knowing of philosophy is Absolute Freedom’s knowledge of itself. Or on the practical side, insofar as Absolute Freedom is will, Absolute Freedom’s self-willing and selfdetermination.55 Since in free thinking the knowing of the Absolute cannot be differentiated from the being of the Absolute, the knowing of the Absolute in free thinking constitutes the very being of the Absolute. Accordingly, the Absolute is constituted by its self-knowledge.56 Since this self-knowledge is only achieved by the overcoming of knowledge and the achievement of non-knowing, the self-knowledge of the Absolute must be construed as a Nicht-wissendes Wissen.57 Absolute Knowing undergoes a three-fold movement: Nicht-Wissen transforms itself into Wissen, which in turn returns back to its origin, NichtWissen, out of its self-alienation. Likewise, Absolute Being undergoes a parallel three-fold movement: undifferentiated Being transforms into differentiated beings, which in turn returns back to its origin, undifferentiated being, out of its self-alienation.58 The whole movement of philosophy is contingent upon holding the Absolute Subject separate from our knowledge,59 for the union of philosophical thinking and the Absolute Subject is only achievable in the recovery and rebirth of Nicht-Wissen. As Schelling teaches us in

Free Thinking in Schelling’s Erlangen Lectures  99 the Freedom Essay: “The Understanding is born in the genuine sense from that which is without Understanding” and “without this preceding darkness creatures have no reality. Darkness is their necessary inheritance.”60 Just as in Hegel’s thought, in Erlangen the movement of reason and philosophy consists in three moments, in which the end is the return of the beginning. Unlike in Hegel’s thought, knowledge is relegated to the second moment of the triad, and does not go beyond the negative relation of the Absolute to itself, whereas non-knowing is the form of the whole.61 Schelling describes this third moment as “Wiedergebrachte Freihiet” and “weiderhergestellte Anfang.” Free thinking overcomes the crisis that is generated by Wissen by going beyond Wissen. The free-thinking of reason heals the division without reducing all differences to one identical unity.62

Conclusion By providing his outline for a theory of philosophy in On the Nature of Philosophy as Science, Schelling develops a meta-metaphysics in which he takes the activity and structure of philosophy as a metaphysical and epistemological inquiry as the object of his inquiry. Because meta-metaphysics belongs to metaphysics, it appears that Schelling’s outline for a theory of philosophy and metaphysics in On the Nature of Philosophy as Science must undermine itself. The reason is simple: since metaphysics is self-destructive, and meta-metaphysical reasoning furnishes a metaphysical theory, it follows that meta-metaphysics must also be self-destructive. Meta-metaphysics cannot escape the fate of metaphysics. If we consider that in Erlangen Schelling has construed the various ways that knowing is related to Being, and has thereby illuminated the structure of Being vis-à-vis knowing, Schelling has made claims about the structure of Being in the essay. Indeed, by making the claim that the Absolute eschews knowing, one is not only making a claim about knowing, but about the structure of Being. Accordingly, Schelling’s own meta-metaphysical claims about metaphysics and philosophy must also be self-destructive. By means of free thinking, metaphysical knowledge destroys itself. Through ecstatic self-destruction, the Absolute becomes manifest as that which transcends knowing. Free thinking is that through which the intellect intuits the Absolute; it is that through which the very existence of the Absolute is made manifest. Since free thinking destroys the very metaphysical knowledge that free thinking posits about the Absolute in order to reveal what the Absolute is, pointing out the self-destructive character of knowledge in free thinking cannot refute Schelling’s position. Rather, pointing out the self-destructive character of knowledge in free thinking only serves to further justify Schelling’s position and embolden it. Conceived in this way, the fact that philosophy is self-destructive no longer entails that philosophy ought to be abandoned. Instead, what is necessary is for the philosopher to adjust her expectations. Rather than expect

100  Gregory S. Moss philosophy to successfully correlate her concepts to the Absolute, the task of philosophy must be re-envisioned as an activity whereby the incomprehensible is revealed. In this way, philosophy cannot be deemed a failure if it fails to conceptually articulate the whole, for its goal is not to conceptually articulate or achieve demonstrative knowledge of the Absolute, but to disclose the incomprehensible by virtue of its very self-destruction. When I follow Schelling into Absolute Freedom by means of free thought, I no longer expect a fully articulated system as the result; instead I expect to intuit the Absolute in ecstasy. Schelling calls us to re-think philosophy as a practice, rather than as a scientific doctrine. On the whole, it is hard to miss the explicitly mystical formula that Schelling has developed in Erlangen: in order to know the Absolute, which is beyond Wissen, one must detach from the self, and thereby become unified with the Absolute in an immediate experience of the God beyond God. But rather than abandon philosophy to mystical insight, Schelling has reconceived philosophical method as the practice of free thinking so that philosophy becomes not only compatible with, but necessary for the mystical insight into the Ewige Magie.63 Free ecstatic thinking offers a systematic, though not scientific, method for mysticism and thereby enriches the mystical tradition. In his reflections in Erlangen on free thinking, one finds an early and uniquely mystical and paradoxical expression of the difference between negative and positive philosophy. This is the case despite the fact that Schelling certainly seems to have later abandoned the highly paradoxical and strongly mystical formulation of positive and negative philosophy in his later thought, e.g., The Grounding of Positive Philosophy. In Erlangen, negative philosophy is constituted by a kind of thinking that attempts to correlate concepts with their proper objects. Here positive philosophy, in contrast, transcends determinate conceptual structures by means of free thinking. In other words, positive philosophy has the pure Daß, the that, not the Was, or what, as its object.64 In Schelling’s Erlangen lectures, the positive significance of indeterminacy as the object of free thinking is inseparable from his conception of positive philosophy. The Absolute breaks through all forms of knowing, and this truth itself cannot be construed by any formula without the Absolute breaking free. Schelling’s account of philosophy in On the Nature of Philosophy as Science can only be successful if his account is self-destructive. Schelling himself does not appear to acknowledge this consequence in his lecture. Nonetheless— Schelling’s meta-metaphysics can only successfully reveal what the infinite is in itself if it also self-destructs. In other words, if Schelling’s account succeeds in the usual way that we expect it to, and does not self-destruct, then it cannot reveal the infinite as that which cannot be confined to any formula whatsoever. Hence, in order for Schelling’s text to reveal the Absolute, it must destroy itself, and in virtue of experiencing that self-destruction, we too can glimpse the indeterminate Übergottheit that no formula can

Free Thinking in Schelling’s Erlangen Lectures  101 fully capture. The Absolute must exceed Schelling’s formula of the Absolute. Ironically, if we insist that Schelling’s account of metaphysics is an exception to the principle that metaphysical accounts are self-destructive, then Schelling’s account cannot be consistent. It can only be consistent if it negates itself, if it is not consistent. Schelling reveals to us the indeterminate Übergottheit by exploding his own text in mystical irony. His philosophy does not say the Absolute, but it shows it by performing what it says. By giving himself up in Selbstaufgegebenheit, Schelling escstatically goes beyond himself into the Absolute. He performs the Ekstasis that is constitutive of the ασυστασία, the want of cohesion of which he speaks. The question is not whether Schelling has gone out into the Spannungslosikeit, but whether we want to accompany him along the way. Schelling errs, but it is a necessary erring, an erring that goes out of itself and into das freie Denken.

Notes 1 Gabriel (2013, p. 12). 2 Ibid., p. 13. 3 Schelling (1927a, p. 215). All future references to this text will be cited as ‘SW’. 4 SW, p. 215. 5 Ibid. 6 Ibid. 7 This is a common theme throughout Schelling’s writings throughout his life. For instance, philosophy of nature and transcendental philosophy constitute two sides of the Absolute in Schelling’s earlier idealistic phase. See Schelling (2001, p. 6). 8 SW, p. 215. 9 Ibid., p. 217. 10 Ibid. 11 Ibid. 12 Ibid. 13 Ibid., p. 216. 14 Ibid., p. 13. 15 Ibid., p. 14. 16 Ibid., p. 235 (Gabriel’s translation: Gabriel [2013, pp. 14–15]). 17 SW, p. 219. 18 Ibid., p. 216. 19 Ibid., p. 210. 20 Ibid., p. 210. 21 Ibid., p. 212. 22 Gabriel (2013, p. 15). 23 SW, p. 211. 24 Ibid., p. 210. 25 Ibid., pp. 212–213. 26 Ibid., p. 219. 27 (This is Gabriel’s translation.) See Gabriel (2013, p. 18). 28 SW, p. 225. 29 Ibid., p. 226. 30 Ibid., p. 219. 31 Ibid., p. 220. 32 Ibid., p. 227.

102  Gregory S. Moss 3 Ibid., p. 224; Gabriel (2013, p. 19). 3 34 SW, p. 224. 35 Ibid., p. 230. 36 Ibid., p. 238. 37 Ibid., p. 241. 38 Ibid., p. 228. 39 Ibid., p. 229. 40 Ibid., pp. 235 and 228. 41 Ibid., p. 235. 42 Ibid., pp. 239–240. 43 For my own exposition of philosophical method, see my book chapter “Negative Realism.” Moss (forthcoming in Das Problem Des Anfangs, edited by Christoph Asmuth and Jesper Rasmussen). There I develop a very similar methodology, though there I motivate the methodology by appealing to self-reference and explicitly recognize the necessity of affirming the contradiction. 44 SW, p. 243. 45 Ibid., p. 242. 46 Ibid., pp. 236–237. 47 Ibid., p. 218. 48 Ibid., p. 239. 49 Ibid., pp. 217–218. 50 Ibid., p. 217. 51 Ibid., p. 246. 52 Ibid., p. 226. 53 Ibid., pp. 229–230. 54 Ibid., p. 233. Schelling also speaks of the human being as “God-positing consciousness.” See Gabriel (2013, p. 95). 55 SW, p. 227. 56 To put it in Hegelian terminology, the Absolute self-knowledge attained in ecstasy is Schelling’s version of Absolute Spirit. 57 SW, p. 233. 58 Ibid., pp. 222–223. With this complete description of the process by which philosophy returns to itself out of its self-alienation, Schelling provides us with his “Grundriss einer eigentlichen Theorie der Philosophie.” Philosophy is selfknowledge. Meta-metaphysics is about itself: meta-meta-metaphysics takes philosophy, whose task primarily consists in self-knowledge, as its object. Although here in Erlangen the three-fold movement begins and ends with the Absolute, Schelling appears later to have modified his method. See Schelling (2007, p. 175). 59 SW, p. 233. 60 Schelling (2006, p. 29). 61 Hegel himself identifies his system of philosophy, in particular the logic of the concept, with the Holy Trinity. Hegel correlated each person of the Trinity with a moment of the concept and one domain of his system of philosophy: “Gott in seiner Allgemeinheit ist das Logische, Gott in seiner Besonderheit ist die Natur, Gott in seiner Einzelheit ist der Geist.” See Hegel (1992, pp. 145–146). 62 SW, p. 76. 63 Gabriel is right to point out that Schelling rejects Quietism in the 1840s in his twelfth lecture in (Schelling 1927b), Philosophische Einleitung in die Philosophie der Mythologie (SW, XI, p. 277). However, not all forms of mysticism are created equal. Although Schelling rejects both mysticism and quietism in his later works, it is undeniable that in Erlangen Schelling profoundly appropriates and offers a methodologically enriched form of hermetic mysticism in the vein of Jakob Böhme, who exercised a profound influence on Schelling. This form of hermetic mysticism in which the Absolute knows itself by means of free

Free Thinking in Schelling’s Erlangen Lectures  103 philosophical thought is profoundly different from the Jacobian form of mysticism that Schelling rejects quite early in his dispute with Jacobi. Still, in Erlangen there are interesting similarities between Schelling’s emphasis on ecstasy and Jacobi’s concept of rapture. For example, Jacobi writes: “We have to make use of the expression “intuition of reason” because language does not possess any other way to signify how something that the senses cannot reach is given to the Understanding in feelings of rapture and yet given as something truly objective, and not merely imaginary” Jacobi (2016, p. 563). Some twenty years or so after giving these lectures in Erlangen, Schelling has clearly moved away from formulating his positive philosophy in terms of hermetic mysticism, as is clear from the Grounding of Positive Philosophy (1842). Here he states that positive philosophy is not to be identified with mysticism or theosophy, and is a “new creation.” Nonetheless, even at this late date Schelling is clear that mysticism “at no time, not even now, has been overcome” and continues to have high praise for J. Böhme. By 1842, Schelling’s positive philosophy does not appear to stand in stark contrast to science. By 1842, Schelling writes that positive philosophy aims to develop a scientific method, whereas mysticism is unscientific. See Schelling (2007, pp. 174–176). In order to transform positive philosophy into science, it would make sense for the later Schelling to turn away from the radical selfdestruction of science that is so pronounced in his earlier Erlangen lectures. 64 Schalow (2001, p. 115). I would like to thank Alexander Bilda for thoughtful insights on how Schelling’s later philosophy diverges from the more paradoxical and mystical formulations of his Erlangen lectures.

Works Cited Hegel, G. W. F. 1992. Vorlesung Über Logik und Metaphysik, Heidelberg 1817. Mitgeschrieben von F. A. Good, edited by Karen Gloy. Hamburg: Meiner. Gabriel, Markus. 2013. Transcendental Ontology. London: Bloomsbury. Jacobi, Friedrich Heinrich. 2016. David Hume. In The Main Philosophical Writings and the Novel Allwill, translated by Georgie Di Giovanni. Montreal: McGillQueen’s UP. Moss, Gregory S. forthcoming. “Negative Realism,” in The Problem des Anfangs, edited by Christoph Asmuth and Jesper Rasmussen. Würzburg: Königshausen & Neumann. Schalow, Frank. 2001. Heidegger and the Quest for the Sacred: From Thought to the Sanctuary of Faith. Dordrecht: Springer. Schelling, F. W. J. 1927a. “Über die Philosophie als Wissenschaft,” in Schellings Werke, Fünfter Hauptband, edited by Manfred Schröter. München: D.H. Beck’sche Buchhandlung. ———. 1927b. “Philosophische Einleitung in die Philosophie der Mythologie,” in Schellings Werke, Fünfter Hauptband, edited by Manfred Schröter. München: D.H. Beck’sche Buchhandlung. ———. 2006. Philosophical Investigations into the Essence of Human Freedom, translated by Jeff Love and Johannes Schmidt. Albany: SUNY Press. ———. 2001. System of Transcendental Idealism (1800), translated by Peter Heath. Charlottesville, VA: University Press of Virginia. ———. 2007. The Grounding of Positive Philosophy, translated by Bruce Matthews, Albany: SUNY Press.

4 Indeterminacy, Modality, Dialectics Hegel on the Possibility Not to Be Nahum Brown

Introduction Hegel’s relentless investigation into the nature of negativity has been well documented in countless commentaries about Hegelian logic and Hegelian thinking in general. Jean-Luc Nancy’s The Restlessness of the Negative, Slavoj Žižek’s Tarrying with the Negative, and Karin de Boer’s the Sway of the Negative, name some of the most prominent book-length studies of Hegel’s preoccupation with negativity, negation, and dialectics. One of Hegel’s most important insights about the nature of modal reality also begins from a preoccupation with negativity. As the basis of the “Actuality” chapter of the Science of Logic (book 2, section 3, chapter 2),1 Hegel claims that if something is possible, this means that it can become actual, but that it also can not become actual. There are, in effect, two sides of possibility in Hegel’s account. There is the positive side, that something possible can become actual. But there is also the negative side, that something can remain merely possible as the possibility not to be. In this chapter, I argue that the negative side of possibility—that what is possible can also not be—plays a significant role in Hegel’s modal argument, and that, based on this conceptual analysis, Hegel’s commitments to actuality-primacy are more complicated than one might think. Possibility has been traditionally conceived of non-dialectically as the possibility to be. We typically define possibility in terms of “truth verification” as the minimum condition for the validity of a proposition, as when we say that if p is possible, there would be nothing contradictory about p becoming actual. This definition of the possibility to be can also be expanded to apply to things and events, and not only to propositions. To have the possibility to do something—for example, to be able to play an instrument or to travel the world—is usually recognized as a clear indication that, putting aside the question of whether the possibility is probable or not, its actualization can in any event come about, since there is nothing inherently contradictory about it. Possibility serves the function of truth verification in these ways. To recognize that something is possible is to recognize that its actualization is possible. But what is missing from this traditional, non-dialectical

Indeterminacy, Modality, Dialectics  105 conception of modality is that when we recognize something as possible, we imply, at the same time, that what can become actual can also not become actual. Hegel’s dialectical account of possibility exposes this conception not only of possibility’s positive side—to be able to play the flute—to be able to travel the world—but also of possibility’s negative side—that in being able to play the flute, we can also not play it—that in being able to travel the world, we can also not travel it.2 If the negative side of possibility were not also present in this category, possibility would be the same as necessity in the sense that whatever is deemed to be possible would have no other recourse than to become actual. But to uphold this most basic distinction between possibility, as what can be, and necessity, as what must be, is to acknowledge a conception of the negative side of possibility. The difference between the can and the must is that the can includes the can not. Essential to the basic conceptual terrain of the possibility not to be is the difference between “being able not to be” and “not being able to be.” That something can not happen is conceptually quite different from the notion that something cannot happen.3 Hegel’s treatment of the possibility not to be can be traced back to the Metaphysics 9.3, to Aristotle’s claim, against the Megarian School of Actualism, that even when a builder is not in the act of building, the builder does not lose the capacity to build but retains this as proof of the existence of unactualized potentiality.4 Hegel’s emphasis on the negative side of possibility, along with Aristotle’s pioneering work on this theme, comes to be a significant precursor for many contemporary continental discussions of modality that attempt in one way or another to make use of the possibility not to be, from Martin Heidegger’s prioritizing of possibility over actuality in Being and Time5 to phenomenological accounts of possibility that place existence ahead of essence, from Agamben’s work on the political ramifications of the potentiality not to be in Homo Sacer6 to Bernard Stiegler’s premonition in the Technics and Time series that our contemporary technological age suffers from the malaise of over-abundant possibilities.7 This legacy of possibility and negativity from Aristotle and Hegel to the contemporary continental tradition is an alternative to mainstream theories of modal reality, which begin primarily from a definition of possibility as truth verification, as with modal logic and possible world semantics, which marginalize the negative side of possibility by emphasizing the role that possibility plays for the sake of actualization. Even Aristotle’s acknowledgement of the ontological status of the potential not to be in his defense against the Megarians can be interpreted as an aberration to his more influential arguments in the Metaphysics 9.8, where he states that actuality is prior and more primary than potentiality.8 Similarly, Hegel’s preoccupation with negativity and his conclusions about “absolute possibility” (SL 486, WL 213) and “absolute contingency” (SL 488, WL 216–217) have, more often than not, been covered over by the entrenched myth that Hegel’s investigations into the nature of negativity serve his ulterior motive to subsume difference

106  Nahum Brown under the dominance of identity and to subsume possibility under the hegemony of actuality. The aim of this chapter is to make sense of Hegel’s very complicated, far from obvious, commitments to the ontological status of the possibility not to be, and to situate his insights as foundational for Heidegger, Agamben, and others. Hegel’s commitments begin from his celebration of the thesis that possibility is as much about non-actuality as it is about actuality. Based on these commitments, Hegel then explores the category of the possibility not to be by focusing on the problem of how to actualize both sides of possibility as one actuality. This leads straight away, for Hegel, to a modal iteration of contradiction. What is fundamentally impossible, and therefore contradictory, would be to actualize the possible itself. However, from this impasse, Hegel deduces the rest of the modal categories of reality as an expansion of actuality that comes to include the negative side of possibility as part of the process of actualization. The following six section headings of this chapter outline Hegel’s analysis of the possibility not to be. The first section, “Hegel’s ‘Actuality’ Chapter,” gives a brief summary of where the “Actuality” chapter stands in relation to the Logic and presents a defense against the objection that local themes of the Logic should not be treated as stand-alone arguments. The second section, “The Two Sides of Possibility,” discusses Hegel’s acknowledgement of the traditional conception of possibility as truth verification and as the possibility to be. In contradistinction, this section also examines Hegel’s statements about the possibility not to be in formal terms. I then explore in the third section, “Possibility and Contradiction,” what I think is Hegel’s groundbreaking and controversial conclusion about modality: that actuality must be able to express both sides of possibility at once. While the actuality of both sides of possibility leads, at first glance, to an unresolvable modal paradox, that is, to the contradiction of conjoining A and -A, the fourth section, “the Actuality of the Possible Itself,” anticipates why Hegel nevertheless commits us to the seemingly untenable actualization of both sides of possibility. The fifth section, “Contradiction and Contingency,” explores Hegel’s formal resolution of this paradox. It also briefly points to further resolutions in the “Real” subchapter as the actuality of conditions and the “Absolute” subchapter as the actuality of substance. In the final section of this chapter, “Possibility and Totality,” I discuss Hegel’s vision of the actuality of the possible itself in terms of his larger project to reconceive of totality as both the positive and the negative together.

Hegel’s “Actuality” Chapter The Greater Logic contains a daunting array of themes, with hundreds of complex arguments, written in what might seem to be an overly dense conceptual language. Hegel divides the book in two different ways. He divides it into two volumes: into an objective and subjective logic. But he also divides

Indeterminacy, Modality, Dialectics  107 it into three doctrines: into the Doctrine of Being, the Doctrine of Essence, and the Doctrine of the Concept. Hegel further divides the Doctrine of Being (first published in 1812) into “Quality,” “Quantity,” and “Measure”; the Doctrine of Essence (first published in 1813) into “Essence as Reflection Within,” “Appearance,” and “Actuality”; and the Doctrine of the Concept (first published in 1816) into “Subjectivity,” “Objectivity,” and “the Idea.” One common way to read the book is to view it as a full-scale deduction of the fundamental categories of thought and reality.9 Much like Kant’s Critique of Pure Reason (A edition, 1781; B edition, 1787), the Logic aims to expose universal categories; however, contrary to Kant, Hegel claims to uncover these categories (arguably hundreds of them) presuppositionlessly, dialectically, and developmentally, rather than to deduce all of the categories (in Kant’s case, twelve categories) as one exhaustive list of axioms. The categories of the Logic range from the simple structures of immediacy, the initial determinations of being, of quality, limit, and finitude, as well as the basic determinations of quantity and measure such as indifference, standing-beside-one-another, number, and arithmetic, to the more complex determinations of essence, where a thing has properties and can maintain a consistent identity throughout various fluctuations; relationships of reflection, identity, difference, contradiction, appearance, existence, part and whole, possibility and actuality, cause and effect, and reciprocity. We can view the Doctrine of Being as an exposition of the most basic categories of being, as what is generated from an investigation into that which being fundamentally is. We can view the Doctrine of Essence, in contrast, as an exposition of the more complex categories that are produced from the question of what being is, rather than from the surface-level description of that which it is. Essence emerges as the substratum of being, both as what being fundamentally is in essence, but also as the alienation of being from itself. We can then view the Doctrine of the Concept as an exposition of those categories that form from the reconciliation of being and essence together, as the resolution of this alienation, as the recognition that the essence of being is nothing other than being’s own self-revelation. The chapter “Actuality” appears as the penultimate chapter of the Doctrine of Essence under the section with the same name, “Actuality.” It comes after the “Concrete Existence” and “Appearance” chapters, before the “Substantiality,” “Causality,” and “Reciprocity” subchapters of “the Absolute Relation,” and is situated alongside a revealing remark on Spinoza and Leibniz. “Actuality” belongs to the Doctrine of Essence because what is actual has emerged from a prior source in possibility. This source is the modality of essence, that what is as simple immediate being has nevertheless come forth into actuality from the negativity of the possible. Although for this reason the “Actuality” chapter belongs to the Doctrine of Essence, this chapter is also an intermediary chapter between essence and the concept. Because actuality maintains the contraries of itself in possibility, it is both itself and the opposite of itself, but it is this as itself, that is, as the concept.

108  Nahum Brown Hegel further divides the “Actuality” chapter into three distinct types of modality: into formal, real, and absolute modality. Each type of modality leads to a network of modal themes and definitions surrounding possibility, actuality, necessity, and contingency that are too complex to address in detail here.10 To summarize briefly, on my reading, formal modality (SL 478–481, WL 202–207) explains Hegel’s initial account of logical possibility, as well as anticipates how Hegel incorporates a traditional definition of possibility as whatever is non-contradictory into his own argument, while ultimately demonstrating why he views this definition as an assumption that must be overcome. Real modality (SL 482–485, WL 207–213) explores the overcoming of this traditional definition of possibility to a limited and implicit extent by outlining a modal interpretation of context-related possibilities, which reintegrate unactualized possibilities through conditional actualizations. The real modality passages of the chapter present Hegel’s theory of dispersed actuality, where the possibilities of one thing are contained in the actualities of others, as well as his theory of conditional actualization, where something initial becomes actual through a series of possibilities, and also his theory of relative necessity, where possibilities are embedded in other actuals and must be drawn out, but gain from this otherwise inaccessible formations of determinate content. The third type of modality—absolute modality (SL 485–488, WL 213–217)—is then the explicit culmination of Hegel’s dialectical account of modality. Absolute modality demonstrates why substance results from real modality in the first place, how to think of a many-substance system as embedded within a one-substance system, and why the absolute necessity of total inclusion invokes contingency as the final consequence of the chapter. Because of the book’s complexity, and because Hegel presents the Logic as one complete, systematic, and developmental account of thought and reality, with the implication that not any one division of the book should take precedence over any other, commentators of Hegel have largely objected to the strategy of isolating any specific argument or theme of the Logic, proposing instead that interpretations should appeal to the broadest consequences of Hegelian thinking generally. To isolate one chapter of the Logic, in this case “Actuality,” inevitably raises questions about whether Hegel’s arguments can be discussed out of the context of the book’s overarching trajectory. What might appear to be even worse would be to isolate a specific theme within a specific chapter, in this case the theme of the possibility not to be. Commentators who have this objection argue that isolating Hegel’s arguments threatens to dislocate the themes of the book from the book’s systematic purpose. The prevalence of this objection has left many of Hegel’s local arguments under-analyzed. Houlgate articulates this objection quite well and also offers a strong defense against it in the “Introduction” to his book The Opening of Hegel’s Logic. “It is clearly very tempting when approaching Hegel,” Houlgate writes, “to think that the whole picture is actually of primary importance and

Indeterminacy, Modality, Dialectics  109 that the details of individual arguments are secondary or even incidental— mere ‘moments’ of a totality that constitutes the real truth or mere ‘examples’ of some universal, omnipotent dialectical principle.”11 Houlgate’s primary defense against this criticism, which he uses to establish his project of analyzing the opening movements of the Logic, is to point out that there is a double-standard: When we read Descartes’ Meditations and Spinoza’s Ethics, we are urged to weigh individual arguments very carefully and consider whether or not they are valid. But when it comes to Hegel, the main point in the eyes of many seems to be to get a rough sense of the whole forest and not to worry too much about the trustworthiness of the individual trees. . . . If we are to follow Hegel himself [however] . . . the properly philosophical way to approach his texts is not to look in the prefaces and introductions for intimations of his general conception of dialectic or spirit, but to look in the main body of his texts at the many particular analyses.12 While there are certainly disadvantages to isolating one theme from one chapter of the Logic, I generally agree with Houlgate that scholarship on Hegel also needs to be able to analyze local arguments, and that we lose track of some of the most important insights of Hegel’s thought if we limit ourselves to only the broadest statements about Hegelian philosophy. The concept of the possibility not to be in the “Actuality” chapter is one such local argument from Hegel that needs to be treated in detail. After all, the “Actuality” chapter is Hegel’s version of Aristotle’s momentous Metaphysics book Theta (book 9). It deserves to be scrutinized and dissected with the same care that centuries of commentators from Aquinas13 to Heidegger14 have given to Aristotle’s stand-alone arguments from modality. What follows, then, is a brief analysis of the motif of the possibility not to be in Hegel’s modal argument, with the aim of contributing to a growing body of textual and critical analysis about the “Actuality” chapter, which can be situated from within the complex mechanics of the Logic and of Hegelian thinking as a whole, but can also be treated as a stand-alone argument.

The Two Sides of Possibility Hegel’s analysis of the possibility not to be begins from two seemingly innocuous claims, (1) “actuality is . . . concrete existence” and (2) “what is actual is possible” (SL 478, WL 202). We make a simple inference from the immediate, self-evident existence of actuality to its status as possibility. If I am reading the newspaper, then reading the newspaper must have been possible and cannot have been impossible. Immediate actuality has this authority. Its possibility is obvious and cannot be contested. This is why we

110  Nahum Brown attribute a high level of truth to the immediate presence of what is already there. By recognizing that whatever is actual is possible, we therefore recognize possibility as truth verification. What is actual is a fact of existence, but it is nevertheless the role of possibility to verify this. In this way, Hegel anticipates the traditionally analytic definition of possibility as the minimum condition for the validity of a proposition. He also thereby anticipates the traditional mainstream conception of possibility as the possibility to be actual, in the sense that if something is already actual, it obviously can become actual. But it is equally apparent from the inference of these two claims—(1) “actuality is . . . concrete existence” and (2) “what is actual is possible”— that the possible is not necessarily actual. Actuality entails possibility, but possibility does not in the same manner entail actuality. This is one of the basic points Fitting and Mendelsohn make in their book First-Order Modal Logic: “Now, p ⊃ ◊p (i.e., It’s actual, so it’s possible) is usually considered to be valid . . . but its converse, ◊p ⊃ p (i.e., It’s possible, so it’s actual) is not [valid].”15 While the statement “it rains” necessitates “it possibly rains,” the statement “it possibly rains” does not necessitate “it rains.” This is the case not only for truth functional sentences, but for entities and things in the world, as well. Although I might buy a house next week, the possibility of this does not necessitate its actuality. This is even more obvious when we think of examples of what is merely possible, where the possibility as such has never been demonstrated in actuality. Fantasy and science fiction stories offer countless examples of what is merely possible, of what has not yet and might not ever become actual. While we can speculate about these various possibilities, their mere projection in possibility does not necessitate their actuality. In this positive sense, possibility is merely the reflection of actuality into itself. “ ‘A is possible,’ ” Hegel writes, “says no more than A is A” (SL 479, WL 203).16 If the sea battle is possible, then it would not break the logical coherence of the event if it were to become actual. In this sense, the possible is already predisposed to become actual. There is no difference added to the content of the possible when it becomes actual. But to say that the sea battle is possible is only to say that if it were to become actual, this would not be impossible—which is to say nothing at all. To say “A is possible” is to express only the most empty of determinations, that what is possible can be, because if it were to become actual, this would be no different than this content as possibility.17 However, Hegel acknowledges not only the positive side of possibility as the tautological validity of actuality when it is reflected into itself, but equally the negative side of possibility. The possibility of actuality turns out also to posit the opposite of the actuality. Hegel writes: “A is A”; then, too, “-A is -A.” These two statements each express the possibility of . . . content determination. But, as identical statements,

Indeterminacy, Modality, Dialectics  111 they are indifferent to each other; that the other is also added, is not posited in either. Possibility is the connection comparing the two; as a reflection of the totality, it implies that the opposite also is possible. It is therefore the ground for drawing the connection that, because A equals A, -A also equals -A; entailed in the possible A there is also the possible not-A.” (SL 480, WL 204) Hegel claims that if A is possible, this means not only that A is identical with itself (that A is A), in the truth-functional sense of the possibility to be actual, but also that the opposite of its identity (-A) is equally possible, in the sense of the possibility not to be. As the most general projection of content, possibility is the relating ground between these two equally true yet empty articulations of the law of identity, that A is A but equally that -A is -A. All possibilities of content are contained within these two articulations of the law of identity. Possibility is therefore the totality of form. It contains both A and -A together. Since it expresses both sides, it exhausts all of the permutations of actuality whatsoever (i.e., there is no possibility for A other than both A and -A).

Possibility and Contradiction Hegel then writes about these two sides of possibility that: As such [possibility] is the relationless, indeterminate receptacle of everything in general.—In this formal sense of possibility, everything is possible that does not contradict itself; the realm of possibility is therefore limitless manifoldness. (SL 479, WL 203)18 When Hegel refers to possibility as an “indeterminate receptacle,” he invokes both sides of possibility at once. Possibility is merely the same content as the actual—nothing more than this—and yet it is also the opposite of the actual. One major insight from Hegel about modality lies in this subtle and complicated difference. On the one hand, possibility is merely the same as the actuality. Yet, on the other hand, possibility contains the opposite of the actuality. There is no additional content, and yet the opposite is posited as well. Indeterminacy is produced from this conception of possibility that is not different from the actuality but is also the opposite of the actuality as well. Because the possible contains both the actual and the opposite of the actual, possibility is boundless, indeterminate multiplicity (die grenzenlose Mannigfaltigkeit). As the unity of being and nonbeing, possibility is a relationless, indeterminate receptacle, open to the being, nothing, and becoming of anything and everything, of any content whatsoever. In this way, Hegel defines possibility formally through the

112  Nahum Brown law of non-contradiction. Propositions and entities are formally possible insofar as they are not contradictory. This makes possibility boundless. Anything and everything is formally possible so long as it does not contradict itself.19 But every manifold is determined in itself and as against an other: it possesses negation within. Indifferent diversity passes over as such into opposition; but opposition is contradiction. Therefore, everything is just as much contradictory and hence impossible. (SL 479, WL 203)20 Hegel then finds from this same initial definition of possibility as indeterminate receptacle the seemingly strange conclusion that whatever is formally possible is just as much something impossible and self-contradictory. Since possibility contains both the identity of the existent-actual and the contrary of the existent-actual, and contains both equally as one unity, everything is possible that does not contradict itself; however, the totality of everything is just as much something impossible and self-contradictory. This is the case because possibility harbors self-contradiction within the function that it serves as actuality’s identity-with-self. This is why Hegel claims provocatively that “possibility is contradiction, or it is impossibility” (SL 479, WL 204). Commentators have offered a great deal of explanation to justify Hegel’s seemingly paradoxical claim that “everything is possible” and “everything is just as much something contradictory and therefore impossible.” The main branch of this comes from John Burbidge, who suggests that Hegel means everything together is impossible, stressing the universality of everything, while at the same time “everything” is possible, stressing the mere possibility of each determinate thing.21 Although both walking and not walking are equally possible—everything is possible in this way—to both walk and not walk in the same time and manner would be impossible. While I think this branch of commentary is plausible, I also think that Hegel actively intends the passage to be paradoxical. If we explain why the passage is not paradoxical, we lose much of what is important about Hegel’s modal argument. In this respect, Lampert really interprets Hegel well when he emphasizes how polemical the passage is, arguing that “the function of a possibility is to express the totality, but that no one possibility can express everything the totality expresses without generating contradictions. Each possibility thus fails to express all that it itself expresses.”22 When we refer to something as possible, we refer to both its positive and negative side. But to actualize both sides of possibility at once would be to actualize contradiction. To have the possibility to play the flute means both to be able to play it and to be able not to play it. However, to actualize both playing and not playing the flute as one activity would be to actualize contradiction. Likewise, I can move to Paris or Seoul. These are

Indeterminacy, Modality, Dialectics  113 two possible directions my life could take. However, to actually move to Paris and Seoul in the same time and manner would be to actualize an impossibility. Since possibility contains both the identity of the existent-actual (A) and the contrary of the existent-actual (-A), everything is possible that does not contradict itself; however, this makes possibility just as much something impossible and self-contradictory. This is the case because possibility harbors self-contradiction (A and -A) within the function that it serves as actuality’s identity-with-self. Hegel emphasizes in this passage that possibility is, at first, only the reflection of actuality into itself, which means that as the minimum condition for the validity of propositions and things, possibility merely affirms the identity of actuality (that if A is possible, then A is A). However, because this simple function of reflection includes the negation of actuality (that what is possible can also not be), possibility therefore also contains the moment of contradiction. When we recall what had seemed to be the innocuous claim from the first lines of Hegel’s chapter, “what is actual is possible,” we are now faced with the paradoxical task of visualizing an actuality of both A and -A. We are now faced with the seemingly contradictory concept of an actuality of the possible itself.23 Now, since [possibility] is determined to be reflective and, as we have just seen, reflectively self-sublating, it is also therefore an immediate and it consequently becomes actuality. (SL 480, WL 204) Possibility has both a positive and negative side, but to actualize both sides would be to actualize contradiction. Try to think of A and -A as one unity. This is impossible. Even the pure conjecture of the imagination cannot visualize A and -A together without transposition. Certainly, I can transpose A and -A. But transposition requires a distinction of time, manner, or place. It requires a disjunction of the actual in the possible. To attempt to render in actuality the immediate status of the possible qua the possible is to present only the indeterminateness, incompleteness, and vagueness of the possible. This is why we fail to grasp the actuality of the possible itself at the formal level. While the possible A contains the possible -A, actuality cannot present this with any determinateness. Or, in other words, its only determinateness is this indeterminacy. Possibility as both the actual and the opposite of the actual turns out to be the first sign of determinateness, according to Hegel. This is why in the “Actuality” chapter Hegel transitions from formal to real modality, from the purely logical stage of possibility as indeterminate receptacle to content-related possibility (i.e., potentiality). Hegel’s rationale for the transition from formal to real modality mirrors, in this way, his rationale at the beginning of the Logic from the indeterminate starting point in being, nothing, and becoming to determinate being. That being transitions

114  Nahum Brown into nothing is proof of its indeterminacy, and yet this is also the first mark of its determinateness as becoming (SL 59–60). Similarly, the actuality of possibility turns out, at first, to be the total absence and indeterminacy of the possible that has no further relation to the actual. And yet this is also the first mark of its determinateness, from which Hegel deduces formal contingency.

The Actuality of the Possible Itself The essential question we face when we explore this textual analysis of Hegel’s passages about possibility is why Hegel would claim that the actual must be able to express both sides of possibility as one actuality. In terms of traditional, non-dialectical conceptions of everyday modality, we normally think of actuality as one of many possibilities but not as the totality of possibility, that is, not as both sides together. While it is possible to live in different cities, if I decide to move to Paris and actually take up life there, I cannot also in the same time, manner, and place, take up life in Seoul or anywhere else in the world. That whole variety of other possible outcomes cannot also come from this decision. The actual is, in this sense, only one of many determinate possibilities. And the possible is that which can be actual but it is also that which can remain unactualized. Why does Hegel claim that thought must also be able to form an actuality of the possible itself, and not only an actuality of one side or the other of the possible, with the other side removed? The primary reason Hegel offers for why thought must actualize both sides of possibility at once is that reality requires the existence of unactualized possibilities as part of the constitution of what makes it true, and that this truth is actuality in the greatest sense of the term. Hegel’s technical argument for this comes from his initial definition of possibility as the reflection of actuality into itself. Possibility functions as both the affirmation of actuality (that if A is possible, then A is A) but also as the opposite of actuality (that if A is possible, then -A is equally possible). “What is actual is,” as Hegel says, “possible.” This seemingly innocuous process of reflection opens a pathway for the contradiction of actuality as both itself and the opposite of itself. When we think about what the seemingly simple statement, “what is actual is possible,” really means, we are forced to admit that actuality ought to express, not only the existent actual that it itself is (e.g., playing the flute), at the expense of its opposite (e.g., not playing it), but also the full expanse of the possible itself, including the possibility not to be. This leads to different types of examples, such as the teleological examples Hegel mentions in the real modality passages, where possibilities are at first dispersed in others and are actualized through conditions. Ultimately, Hegel challenges us to be critical of common-sense assumptions about the nature of modal reality, such as that actualization must always form a disjunction, rather than a conjunction, of the sides of possibility.

Indeterminacy, Modality, Dialectics  115 Hegel offers a unique and controversial modal theory in this one important respect. He not only acknowledges the possibility not to be, which Aristotle and numerous others can be interpreted to have already discovered. Hegel also argues that actualization must be able to take up the contrary sides of possibility as the self-same actuality. In this way, Hegel’s conception diverges from most traditional conceptions of modal reality, which prioritize the ontological status of actualized possibility over unactualized possibility. Whereas most traditional conceptions of modal reality describe actuality as one strand or another of the possible with the other side of the possible removed at the point of actualization (e.g., either playing the flute or not playing it, but not both), Hegel’s theory is unique because he requires actuality to express the totality of possibility. Hegel’s deduction of the modal categories—which continues along from the moment of formal contradiction to formal contingency and necessity, and from this to his teleological account of possibility in the “Real” subchapter, and from this to the absolute modality of substance in the “Absolute” subchapter—depends upon his initial assertion that thought must attempt to actualize both sides of possibility as one actuality. Hegel’s conclusion that the possibility not to be is a significant aspect of modal reality, that, in effect, there are two sides of possibility, and that actuality must be able to express both sides together, exposes, on the one hand, Hegel’s commitment to actuality-primacy, in the sense that Hegel does not let the possible remain completely unactualized. The contradiction of actualizing both sides of possibility arises because thought attempts to actualize, not only the positive side, which is usually conceived to be the normal side to actualize, but also the negative side together with the positive. It is Hegel’s insistence on actuality-primacy, one could say, that leads to the paradox in the first place, where thought tries to form an actuality of the possible qua the possible. But this same conclusion also reveals, on the other hand, Hegel’s commitment to possibility-primacy. For what is actualized is not the divided possible, the actual A at the expense of -A, or vice versa. What is actualized is the totality of possibility (both A and -A together). By claiming that thought must think the actuality of the possible itself, Hegel thereby lifts the ontological status of unactualized possibility to the status of actuality. Traditional, non-dialectical accounts of modality, in contrast, avoid this contradiction by lowering the ontological status of unactualized possibility to non-actuality. Hegel also thereby prioritizes indeterminacy over determinacy. Traditional modal accounts assume that to actualize possibility means to divide it into determinate aspects, where one aspect of possibility becomes actual at the expense of other aspects. But Hegel continuously attempts to leave undivided possibility intact. He thereby begins from the primacy of possibility as indeterminacy. That actualization initially fails to capture both sides of possibility at once marks the first moment of possibility as divided and determinate. However, that thought continuously attempts to present the possible

116  Nahum Brown itself—in what becomes for Hegel “formal contingency”—exposes Hegel’s commitment to the primacy of indeterminacy over determinacy. Determinate possibilities appear, for Hegel, to be derived from and dependent on the more originary actualization of the possible itself.

Contradiction and Contingency Hegel acknowledges that the unmediated actualization of the possible itself leads to an unsustainable contradiction, in which, it turns out, the possible becomes impossible. But this does not stop thought from attempting to conceive of both sides of possibility as one actuality. The immediate, unsustainable contradiction of actualizing both sides of possibility at once generates a series of developmental stages of the modal categories, according to Hegel. Each stage of Hegel’s developmental modality “softens” the contradiction of actualizing the possible qua the possible by expanding our conception of actuality to include the negative side of possibility as part of its constitution. Formally, something cannot both be itself and the opposite of itself since this would lead directly to contradiction. But if the actualization of the opposition inherent in possibility were to become “softened” to the point at which the two sides of possibility were no longer to lead to contradiction, this would affirm the differences that the opposition had contained without causing the erasure of these differences from existence. Essentially, Hegel expands the category of actuality so that it can include the positive and negative sides of possibility as one actuality. On my reading, he expands actuality in three different ways. Hegel calls the first expansion of actuality “formal contingency.” This type of actuality includes the negative side of possibility as what could have been. Hegel then calls the second expansion of actuality “real” or “conditional” actualization. An actuality that is a condition holds both sides of possibility together in the sense that a condition contains the possibilities of other actuals (i.e., the stone is a condition for the statue.) Hegel calls the third type of expanded actuality “absolute actuality.” In the absolute modality passages of the chapter, Hegel anticipates how to conceive of substance as an actuality that is explicitly both sides of possibility together (i.e., both this particular goat and the universal instantiation of a goat). These three expansions of actuality mitigate the contradiction and do not let it fully emerge in its pure, unmediated format. In effect, they dress up the contradiction so that it does not appear as unsustainable. And yet, just as each type of expanded actuality offers an explanation for the possible qua the possible, each likewise posits the possibility not to be as existing alongside and in unity with the actual. To illustrate how Hegel expands actuality to include both sides of possibility at once, I will now discuss in more detail the first type of expanded actuality, “formal contingency,” as a continuation of my textual analysis of the formal passages of Hegel’s chapter. I leave the details of Hegel’s real and absolute account of expanded actuality for other discussions.24

Indeterminacy, Modality, Dialectics  117 Hegel turns from the seemingly unsustainable contradiction of actualizing the possible itself to “formal contingency.” “This unity of possibility and actuality,” Hegel writes, “is contingency.—The contingent is an actual which is at the same time determined as only possible, an actual whose other or opposite equally is” (SL 480, WL 205). In contingency, the negativity of the possible—that what is possible can and can not be—becomes contingent actuality. What is immediately actual is a fact of existence. But this fact is contingent, which means that it is but its other could have been. The negative side of possibility appears through whatever happens immediately to exist in actuality, and in this way, possibility itself is also posited. Because A is a contingent actuality, A posits -A as what equally exists. Contingency thus presents, in a certain respect, the actual A as containing the existence of -A within its own concept. The reason why this is no longer a contradiction is because A and -A contain each other in a relationship of indifference. Let us visualize what Hegel has in mind. The actual world appears in its immediate givenness. This is the simple actuality of concrete existence. The horse, for example, is there in the barn. The other swimmer is there in the lane. If I look at a map of the earth, I see the mountains and the lakes as already there, as what is simply given to their region. Yet, since this that just appears before me is itself possibility, I recognize in what is already there something more than what is already there. This leads to two further points. I recognize alongside the actual-existent that the possibility of its opposite has of itself an existence, and that this existence could have been what is actual. But I also recognize not only that possibility exists, but also that the immediately existing actual only happens to be. While what immediately appears has the authority of truth, it appears at the same time as finite, since it appears with its other alongside it. Hegel is not only saying that in the contingent actual, the other appears too, and that these sides together express the possible itself. He is also saying that actuality depends upon the equal existence of the possibility not to be. The swimmer passes me in the lane; yet, the quality of this simple event is formed in the contingency and instability that what happens could not have happened. I see in the body movement of the other swimmer, not only the appearance of what could have been, but the literal texture of the possibility not to be as it exists in the actuality of what is. Formal contingency shows us that even though what emerges in actualization is only some limited actuality of content, the negative of this actuality equally could have been actual. Formal contingency is itself the immediacy of actuality, yet what is posited in contingency is at the same time the existence of non-actual possibility. In this way, formal contingency initiates a revision of the relationship between existence and possibility because it posits the negative side of possibility alongside the actual. Contingency shows us that the immediate actuality of what merely exists acts as a metaphysical gateway upon the existence of possibility qua possibility. Possibility appears through the limitation of immediate actuality. It appears in what Hegel calls

118  Nahum Brown the empty or superficial sense as the positive identity of the actual. But it also appears in the negative sense, as the possibility of other actuals, and as the possibility of itself qua other.

Possibility and Totality Contingency is for Hegel the first development of an expanded conception of actuality that includes both sides of possibility at once. Contingency is only the first stage of this development, however, because it can only hold the sides of possibility together through the indifferent mode of an actual existent, which could have been otherwise but is not other than it is. Contingency nevertheless initiates a progressive series of modal deductions that witness more and more adequate expansions of actuality, from formal contingency and necessity, to the real modality of conditions, and eventually to the absolute modality of substance. Hegel’s recognition that the two sides of possibility form a totality, and that actuality must be able to express this totality adequately, underlies Hegel’s theory of these modal deductions.25 The term “totality” normally suggests connotations of exhaustion, completion, and finitude. Traditional associations of the totality of possibility present us with the exhaustion of possible outcomes, as when we say, for example, that only certain possibilities can emerge from certain events or that all possibilities are already determined to have a certain outcome. Taken to the extreme, these associations invoke deterministic, theological, necessitarian interpretations of a God who encloses the world in actuality, of a modal reality that leaves little imagination for unactualized possibilities and even less for conceptions of free contingency. At first glance, Hegel’s association of possibility and totality, which we find conspicuously in the first lines of the “Actuality” chapter, might appear to place him firmly among philosophers who privilege actuality and necessity over possibility and contingency. Contrary to traditional connotations, however, Hegel consistently defines totality counter-intuitively as the coincidence of being and negativity, that is, as the complete form A is -A. Hegel’s celebration of the negative side of possibility offers a specifically modal example of a theme that is prominent throughout his work. When he claims famously in the Phenomenology that the “True is the whole”26 or famously in the Logic that being and nothing pass over into becoming (SL 82–83), Hegel has the more general implications in mind of the modal claim that possibility is “the totality of form” (SL 542–543). If we define possibility not only as the positive account of itself in actuality, but also as the negative account as the possibility not to be, possibility then encompasses the totality of any actualization whatsoever in the sense that there is no other permutation of actuality other than that it either is or is not. And yet if this recognition of possibility’s negative side amounts to the exhaustion of all possible outcomes, this is a strange and unprecedented kind of exhaustion since its totalizing effect includes the

Indeterminacy, Modality, Dialectics  119 possible itself, rather than restricts it. To include the negative side of possibility exhausts the form of actualization. Possibly A expresses the totality of actualization because it projects both A and -A. Hegel’s specifically modal claim that “possibility is the totality of form” is a moment of a broader trend in his work to reconceive of totality in terms of negativity and to show how the completion of the form A is -A leads to a fiercely agitated ontology of becoming. This argument appears prominently from the first moment of the Logic. The movement from being to nothing and from nothing to becoming exhibits the utter exhaustion of the most abstract ontological categories as the coincidence of being with its negativity. Because being and nothing pass over into each other, there is no possibility of a remainder or exterior beyond the all-encompassing trajectory of their passage. And yet the reason why becoming is the result of this is because the kind of totality that being and nothing form is productive and inclusive for determinate reality rather than restrictive and exclusive. One might say that the passage from being to nothing, what Hegel calls “ceasing-to-be” in the initial terminology of the Logic, is an expression of Hegel’s reconception of totality in the sense that being encompasses everything in such an abstract way that nothing determinate can exist beyond being. And yet one might also say that the passage from nothing to being, what Hegel calls “coming-to-be” in the initial terminology of the Logic, is also an expression of totality in the sense that nothing, as the negation of being, exhausts all of the permutations that are possible for being. The reason why becoming results from the coincidence of being with its negativity is because the exhaustion of the most abstract ontological categories is at the same time the proliferation in the most agitated and dynamic way of determinate reality. This motif comes up again on almost every page of the Logic, but prominently in “the Essentialities or the Determinations of Reflection” chapter of the Doctrine of Essence. In a remark about the law of contradiction (SL 381–385, WL 74–80), Hegel suggests that the reason why contradiction is more primary than identity is because contradiction contains the totality of the form of reflection, whereas identity contains only one side or the other of this totality. “In contrast to [contradiction],” Hegel writes, “identity is only the determination of the simple immediacy, of inert being, whereas contradiction is the root of all movement and life; it is only in so far as something has a contradiction within it that it moves, is possessed of instinct and activity” (SL 381–382, WL 75). The law of identity (A is A) relates things and propositions only to themselves. The relation of the copula restates the same content throughout. But the sense of totality that comes from contradiction is altogether different. In a contradiction A is -A. The relation between the copula of a contradiction exposes all of the various possibilities of identity by way of the strange unthinkable conjunction that A, by not being itself, is itself. Whereas the identity of A excludes -A, contradiction is a totality in the profound sense that it includes even that which identity excludes, its own non-identity. Contradiction is therefore more primary than identity because

120  Nahum Brown it articulates the totality of form, that A is both A and -A. Contradiction exhausts the formal positions of individuation. This exhaustion is not static or dead being. On the contrary, contradiction interrupts the fixation of identity by exposing what is to its opposite. The interruption of identity is at the same time the totality of identity, since there is no possibility outside of the form that contradiction exhibits for it. All of this suggests that possibility is a further permutation of terms like becoming from the Doctrine of Being and, in the more advanced language of reflection, of contradiction from the Doctrine of Essence. The totality of possibility, or what I have also referred to as the actuality of the possible itself, is the summation of the positive and negative sides of possibility as one aggregate. Together, A and -A expose the totality of possibility not in the sense of determinate limitations, where something can be but other determinate things cannot be, and where all possible determinations are already given. Possibility forms a totality in a different sense than this. It forms a totality as the indeterminate, which precedes the exclusive either/ or disjunction of determinacy. Because possibility projects the alternatives A and -A, it exhausts the form of A. The inclusion of -A, which is the primary significance of possibility for actuality, establishes a totality that is at the same time the opposite of any determinate actuality, and therefore the open engagement of indeterminacy. The history of modal ontology has for the most part obscured the negative side of possibility. Hegel’s argument from modality challenges us to revive this marginalized aspect of possibility. In this way, the “Actuality” chapter offers an outstanding attack on the suppression of the ontological status of the possibility not to be in the history of Western thought about modality. Hegel’s primary contribution to the conceptual analysis of this modal category comes through his exploration of what it would mean to actualize both sides of possibility together. This leads, on his account, to the halting discordance of modal contradiction, where the actualization of the possible itself is conceived of as impossible. But this also leads to Hegel’s signature insight that further modal categories develop from the resilience of forms of actualization that overcome this initial impasse of contradiction. Ultimately, Hegel’s exploration of the actuality of the possible itself reveals a modal aspect of his reconception of totality as the upholding of the indeterminacy that what is possible can also not be. For these reasons, Hegel should be recognized as a precursor for continental modal theorists, such as Heidegger and Agamben, who emphasize the negative side of possibility as constitutive for modal reality.

Notes 1 Hegel (2010, pp. 478–488). Hegel (1969b, pp. 202–217). Hereafter cited as SL/ WL, followed by the English/German pagination. Unless otherwise noted, I have used di Giovanni’s translation of SL and Geraets, Suchting, and Harris’s translation of The Encyclopaedia Logic throughout.

Indeterminacy, Modality, Dialectics  121 2 My analysis of Hegel on the possibility not to be draws heavily from Agamben’s distinction between the can and the can not in “On What We Can Not Do.” See Agamben (2011, pp. 43–45). 3 I take this distinction from Aristotle, who warns us not to conflate can not and cannot, when he explains in De Interpretatione that the opposite of “it is possible for something to be” is not “it is possible for something not to be,” but rather “it is not possible (impossible) for something to be.” If somebody has the capacity to walk, Aristotle points out, this person can choose to walk or not to walk. It would be a mistake to say that when this person chooses not to walk, this person cannot walk. Aristotle (1984, pp. 34–35, 21a34–22a12). 4 Aristotle (1984, pp. 1653–1654, 1046b29–1047b2). 5 “Higher than actuality stands possibility. We can understand phenomenology only by seizing upon it as a possibility.” Heidegger (1962, p. 63). For my related analysis of Heidegger’s modal theory, see Brown (2017a). 6 The “Potentiality and Law” chapter of Agamben’s Homo Sacer offers an ontological-modal basis for the political sovereign paradox in the form of an analysis of Aristotle’s Metaphysics book Theta over the constitutive ambiguity between whether actuality or potentiality is more primary. For my explanation of this argument, see Brown (2013). Besides his essay “On What We Can Not Do,” which contains, in my opinion, one of his most penetrating and concise contributions to conceptual analysis of the possibility not to be (Agamben uses the term “impotentiality”), Agamben also emphasizes the importance of the can not in various related concepts throughout his corpus. This concept appears in The Coming Community as the concept of the “whatever” (1993, p. 9), in Homo Sacer as the relation between example and exception (1998, p. 21–23), and in the “As If” and “Exigency” sections of The Time that Remains as messianic modality (2005, pp. 35–39). This concept also appears prominently in his collection of essays Potentialities (1999). Leland De la Durantaye, an influential commentator of Agamben, goes so far as to call Agamben the philosopher of potentiality. De la Durantaye (2009, p. 4). 7 Stiegler (2010, pp. 202–207). 8 Specifically, Aristotle claims that actuality is prior to and therefore more primary than potentiality in terms of definition (logos), substance (ousia), and in one sense of time (chronos). The only sense in which potentiality precedes actuality is in the sense that the child or seed exists literally, chronologically prior to the adult. Aristotle (1984, pp. 1657–1660, 1049b4–1051a3). 9 For example, see Houlgate (2006, p. 9). 10 I think that the best commentaries of Hegel’s “Actuality” chapter are Henrich (1971), di Giovanni (1980), Houlgate (1995), Lampert (2005), Burbidge (2007), and Longuenesse (2007). 11 Houlgate (2006, p. 4). 12 Ibid., p. 5. 13 See Aquinas (1961). 14 See Heidegger (1995). 15 Fitting and Mendelsohn (1998, p. 5). 16 Hegel probably has Kant’s criticism of the ontological argument in mind when he claims that possibility is the empty reflection of actuality into itself. In chapter three, section four, of the Critique of Pure Reason, “On the Impossibility of an Ontological Proof of God’s Existence,” Kant claims that being is not a predicate and that, in modal terms, there is nothing added to the content when something possible becomes actual. “A hundred actual dollars does not contain the least bit more than a hundred possible ones” Kant (1998). With this statement, Kant criticizes Saint Anselm’s famous inference from the mere possibility of God to God’s

122  Nahum Brown actual existence. For Hegel’s explicit discussion of Kant’s criticism, which appears as part of Hegel’s defense of the transition from being to nothing, see SL 62–66. 17 Hegel’s analysis of the positive side of possibility begins formally in this way as truth verification, but it also carries over into the “Real” subchapter as the determinate, teleological drive of something possible as predisposed to become actual. One of Hegel’s favorite examples of the “real” possibility to be actual appears in the organic life passages of the Logic and the Philosophy of Nature as the seed’s predisposition to become an oak tree. For the sake of simplicity, I will focus primarily on the formal version of Hegel’s argument and leave the real and absolute developments of Hegel’s argument for other discussions. 18 Translation modified. 19 Hegel is obviously critical of this conception of possibility as indeterminate receptacle in the Encyclopaedia Logic when he writes: “[E]ven the most absurd and nonsensical suppositions can be considered possible. It is possible that the moon will fall on the earth this evening, for the moon is a body separate from the earth and therefore can fall downward just as easily as a stone that has been flung into the air; it is possible that the Sultan may become Pope, for he is a human being, and as such he can become a convert to Christianity, and then a priest, and so on . . . Anything for which a ground (or reason) can be specified is possible.” Hegel (1991, p. 216). 20 Translation modified. 21 Burbidge (2007, p. 19). 22 Lampert (2005, p. 75). My analysis of possibility and contradiction has been influenced by Lampert’s reading of the “Actuality” chapter in this respect. 23 I use the phrases “the possible itself” and “possibility qua possibility” interchangeably. These phrases refer to the sides of possibility as one unity, A and -A together, that something possible both can and can not be. The last section of this chapter, “Possibility and Totality,” develops conceptual analysis of this unity, where the actuality of both sides of possibility is expressed as a totality. 24 For my analysis of how expanded actuality applies to the real and absolute passages of Hegel’s modal argument, see Brown, “Possibility Entails Itself in Actuality: Hegel’s Theory of Conditions,” (forthcoming). My analysis of formal contingency also draws from this book chapter. 25 This discussion of Hegel’s reconception of totality draws on my similar discussion about “apophatic totality” in Brown (2017b, 107–129). 26 Hegel (1977, p. 11).

Works Cited Agamben, Giorgio. 1993. The Coming Community. Minneapolis, MN: University of Minnesota Press. ———. 1998. Homo Sacer: Sovereign Power and Bare Life. Stanford, CA: Stanford University Press. ———. 1999. Potentialities: Collected Essays in Philosophy. Stanford, CA: Stanford University Press. ———. 2005. The Time that Remains: A Commentary on The Letter to the Romans. Stanford, CA: Stanford University Press. ———. 2011. Nudities. Stanford, CA: Stanford University Press. Aquinas, Thomas. 1961. Commentary on the Metaphysics of Aristotle, Volume 1. Washington, DC: H. Regnery Company. Aristotle. 1984. The Complete Works of Aristotle: The Revised Oxford Translation, 2 vols. Bollingen Series. Princeton, NJ: Princeton University Press.

Indeterminacy, Modality, Dialectics  123 Brown, Nahum. 2013. “The Modality of Sovereignty: Agamben and the Aporia of Primacy in Aristotle’s Metaphysics Theta.” Mosaic 46(1): 169–182. ———. 2017a. “Aristotle and Heidegger: Potentiality in Excess of Actuality.” Idealistic Studies 46(2): 199–214. ———. 2017b. “Is Hegel an Apophatic Thinker?” in Contemporary Debates in Negative Theology and Philosophy, edited by N. Brown and J. A. Simmons (pp. 107–129). Cham, Switzerland: Palgrave Macmillan. —— Forthcoming. “Possibility Entails Itself in Actuality: Hegel’s Theory of Conditions,” in The Necessity of Freedom in Hegel: Logic, Phenomenology and History, edited by Emilia Angelova. Toronto, ON: University of Toronto Press. Burbidge, John W. 2007. Hegel’s Systematic Contingency. New York, NY: Palgrave Macmillan. De Boer, Karin. 2010. On Hegel: The Sway of the Negative. New York, NY: Palgrave Macmillan. De la Durantaye, Leland. 2009. Giorgio Agamben: A Critical Introduction. Stanford, CA: Stanford University Press. Di Giovanni, George. 1980. “The Category of Contingency in the Hegelian Logic,” in Art and Logic in Hegel’s Philosophy, edited by Warren E. Steinkraus (pp. 179– 200). Atlantic Highlands, NJ: Humanities Press. Fitting, Melvin and Richard L. Mendelsohn. 1998. First-Order Modal Logic. Dordrecht, the Netherlands: Kluwer. Hegel, Georg Wilhelm Friedrich. 1969a. Hegel’s Science of Logic. Amherst, MA: Humanity Books. ———. 1969b. Werke in zwanzig Bänden, 6: Wissenschaft der Logik II. Frankfurt am Main, Germany: Suhrkamp Verlag. ———. 1977. Hegel’s Phenomenology of Spirit. Oxford, UK: Oxford University Press. ———. 1991. The Encyclopaedia Logic. Indianapolis, IN: Hackett Publishing. ———. 2010. The Science of Logic. Cambridge, UK: Cambridge University Press. Heidegger, M. 1962. Being and Time. New York, NY: Harper & Row. ———. 1995. Aristotle’s Metaphysics Θ 1–3. Indianapolis, IN: Indiana University Press. Henrich, Dieter 1971. Hegel im Kontext. Frankfurt, Germany: Suhrkamp. Houlgate, Stephen 1995. “Necessity and Contingency in Hegel’s Science of Logic.” Owl of Minerva 27(1): 37–49. ———. 2006. The Opening of Hegel’s Logic: From Being to Infinity. West Lafayette, IN: Purdue University Press. Hughes, George Edward and Max J. Cresswell. 1996. A New Introduction to Modal Logic. London, UK: Routledge. Kant, Immanuel. 1998. Critique of Pure Reason. Cambridge, UK: Cambridge University Press. Lampert, Jay. 2005. “Hegel on Contingency, or, Fluidity and Multiplicity.” Bulletin of the Hegel Society of Great Britain 51(2): 74–82. Longuenesse, Béatrice. 2007. Hegel’s Critique of Metaphysics. New York, NY: Cambridge University Press. Nancy, Jean-Luc. 2002. Hegel: The Restlessness of the Negative. Minneapolis, MN: University of Minnesota Press. Stiegler, Bernard. 2010. Technics and Time, 3: Cinematic Time and the Question of Malaise. Stanford, CA: Stanford University Press. Žižek, Slavoj. 1993. Tarrying with the Negative. Durham, UK: Duke University Press.

Part II

The Significance of Indeterminacy for Phenomenology, Natural Science, and Ethics

5 Determinable Indeterminacy A Note on the Phenomenology of Horizons Steven G. Crowell

Determinateness, as I shall understand it here, is the property of a thing thanks to which that thing possesses all its other properties in itself, that is, apart from any consideration of the cognitive acts in which those properties come to be known. Even relational properties belong to the thing in this sense: an elephant’s being bigger than a mouse is determinately its property, whether or not there are any mice; a hammer’s being a tool for driving nails belongs to it, even if such a property cannot be defined without reference to other things. Such a view of determinateness underwrites a very widespread form of realism, according to which the aim of cognition is to grasp the determinateness, the truth, of things, and indeterminacy must be understood as a limitation of cognition. Cognition, itself a determinate thing, discovers the indeterminate as the horizon of unknowns in relation to what it knows. But how are we to characterize this horizon? In this note, I take up some aspects of the phenomenological answer in Husserl, Merleau-Ponty, and Heidegger.

Husserl’s Concept of Horizon Husserl’s 1913 book Ideas offers the outlines of a transcendental philosophy grounded in the phenomenological method. Under the heading of “descriptive psychology,” Husserl had already employed the method in the Logical Investigations (1900–1901), but he came to believe that phenomenology, when purified of all psychological interpretation, provided the epistemic ground for all philosophical questions, “including as well all so-called metaphysical questions, insofar as they have possible sense in the first place.”1 At the beginning of Ideas, where the goal is to identify “the entrance-gate of phenomenology”2—the portal to “purifying” phenomenology of all psychological assumptions—Husserl introduced the concept that came to define the phenomenological project: “world.” Here, too, we find his first mention of the “horizon.” Let us listen in on his descriptions which, as he says, “can best be carried out in the first-person singular.”3 No matter what I busy myself with—be it a scientific project of explaining the atomic structure of physical reality or some practical concern for the

128  Steven G. Crowell well-being of myself, my family, or a wider society—“I am conscious of a world endlessly spread out in space, endlessly becoming and having become in time.” This world, and the various things I encounter there, “are simply there for me [. . .] whether or not I am particularly heedful of them.”4 For the scientist who explains trees and rocks as patterns of atoms arranged tree- and rock-wise, the computers and other apparatus upon which her explorations depend are simply “there,” perceived as belonging to the world of what Husserl calls the “natural attitude.” But how should this being there be described? It is not merely that such things are directly perceived. Together with whatever thing I am currently attending to, “other actual objects are there for me as determinate, as more or less well known”—for instance, the whiteboard behind me, on which I have scribbled an equation; the chair upon which I sit while attending to the data on my computer screen; the coffee cup lying in the periphery of my field of vision, ready to be grasped. Thus the focal point of perception is surrounded by a “halo” of what is co-present as “intuitionally clear or obscure, distinct or indistinct.”5 But this is not all, according to Husserl: “What is now perceived and what is more or less clearly co-present and determinate (or at least somewhat determinate), are penetrated and surrounded by an obscurely intended to horizon of indeterminate actuality.”6 The familiarity of the chair I sit on means that it is co-present to me as “somewhat determinate,” but the horizon of my perception stretches out toward what is not perceptually present at all. I can think about such things—remember the route from my house to my campus office, think of its location in Houston, and even imagine the nested contexts that lead out to the “universe.”7 But with each iteration of such expansive “thinking,” what I think becomes ever more indeterminate. As Husserl puts it: “an empty mist of obscure indeterminateness is populated with intuited possibilities or likelihoods”—i.e., motivated imaginings, positings which lack the determinateness of intuited actualities—in which “only the ‘form’ of the world, precisely as ‘the world,’ is pre-delineated” as actual. Thus, Husserl concludes, “my indeterminate surroundings are infinite, the misty and never fully determinable horizon is necessarily there.”8 What can we learn from this description of the natural attitude regarding the theme of indeterminateness? First, that the indeterminacy of the horizon is a necessary feature of perception, and second, that it is never fully determinable. As necessary, the horizon ensures that anything I perceive is found within a context of other things that are co-perceived more or less obscurely, indeterminately, but which can be made determinate through (motivated) experiences that lead from one thing to another. The perceived thing, as Husserl puts it, “leads, of itself, beyond itself” by means of “intentional implications” that link it with things in the horizon in the normative, not causal, manner distinctive of conscious experience.9 But just as necessarily, this sort of determinable indeterminacy eventually gives out: I can

Determinable Indeterminacy  129 never make everything in the horizon explicit in all its properties. Only the “ ‘form’ of the world” holds it all together in actuality. The indeterminateness of the horizon is here contrasted with the “form” of the world. What is the significance of this distinction? In contrast to the changing horizons of my particular experiences, “I find myself [. . .] at all times, and without ever being able to alter the fact, in relation to the world which remains one and the same, though changing with respect to the composition of its contents.”10 The “form” of the world, then, is its unalterable identity as actual. Further, it is not just “my” world but “one and the same world of which we are all conscious, only in different modes.”11 Thus, while I may find that something presenting itself as actual in the world is an “hallucination” or “illusion” and so deny it a place in the actual, the “world is always there as an actuality;” no “doubt” that arises within the world “alters in any respect” the “positing” of the world in the natural attitude.12 But do we not sense a problem here? Is there not some tension between the claim that the horizon of any experience provides for continuous determination of what is indeterminately co-present (i.e., coming better to know what is actually in the world) and the claim that what is found within this horizon is never fully determinable? What is the force of the “never” and of the “fully”? If the “form” of the world guarantees the “unalterable identity” of the actual, is the necessary indeterminateness of the horizon merely a subjective feature of experience? If the world is not itself a thing in the world, does its “unalterable identity” guarantee that the things populating it are themselves fully determinate? And if not, what would continuous determination of such things amount to? As they arise within the natural attitude, these are metaphysical questions. For “pure” or transcendental phenomenology, in contrast, the metaphysical orientation is “bracketed” in favor of a metaphysically neutral investigation into the essential correlation of consciousness and world. The phenomenological reduction or epoché abstains from positing the world as actual (and so also from commitments about the ultimate determinacy or indeterminacy of what is found there) in order to focus on how the sense of actuality is constituted in conscious experience. Such abstention does not deny the world’s actuality but “makes no use of” whatever presupposes it— including the findings of the natural sciences (for instance, questions about quantum indeterminacy)—thematizing, instead, the horizonal character of experience. The controversies surrounding the epoché cannot be explored here.13 Our concern is with the horizon itself and the way Husserlian phenomenology understands the determinable indeterminacy belonging to it—namely, as the starting point of inquiry. On the basis of how things are given, inquiry pursues a path from the relatively indeterminate to the increasingly determinate, or knowledge, guided normatively by a kind of anticipation. From this point of view, the epoché expresses a commitment to reflecting exclusively

130  Steven G. Crowell on what makes up such a path, a commitment which defines the methodological significance of Husserl’s transcendental “idealism.” Let us consider an example. If I am hiking along a path in the high desert, my attention might be drawn to a peculiar rock formation. I need make no judgment to perceive it as an actual rock formation, and neither is any “conceptual activity”14 involved when I perceive it as determined in certain ways—as being a certain color, shape, and texture. But, as Husserl points out, my perceptual intention goes beyond such determinations toward what is still indeterminate about the thing.15 The side of the formation I cannot see is “there” as what is as yet unknown about it. This “unknown” belongs to the horizon of the perception, which involves two dimensions: an internal horizon of the formation itself, the various properties that a motivated exploration can discover in it, and an external horizon, which situates it within a definite place in the world, starting with its place in the landscape.16 The rock formation stands there on the desert floor surrounded by walls of stone towering over it and shading it from the sun, but this cluster also belongs (more or less determinately) to a world in which there are human beings who make things and live their lives in various places, of which this might be (or might once have been) one. From a metaphysical point of view, all of this is just a question of fact: either someone lived here or they did not. Phenomenology, in contrast, puts such facts in abeyance. Only what enters into my current experience belongs to the phenomenon. So what does enter into it? What is the nature of the indeterminateness which belongs to the horizon of my perception? As Husserl notes, the governing perceptual intention that grasps the thing as a rock formation contains “partial intentions” that are “empty” in the sense that what they intend is not sensuously given—such as the side of the formation facing away from me. But these partial intentions, which target the horizon, are not completely empty; the “indeterminacy” of what they aim at is not the “absence of determination” but rather “incomplete determinacy.”17 They are anticipations of what should show up—if the thing really is what I take it to be—when I move around the formation or otherwise explore its perceptual features. In Husserl’s language, the content of any current moment of a perceptual act—its noema, including the noematic meaning and its current “modes of givenness”18—is bound to the content of subsequent perceptual acts of the “same” thing by means of intentional implications. These differ from metaphysical or conceptual implications because they belong to the horizons of my experience, not to the thing considered apart from all experience. Intentional implications are not logical but teleological, in the sense that they chart a path from the “incomplete determinacy” of the perceptual content to its further determination.19 In looking at the rock formation, the back side is “there” as allowing for a certain range of possible views, including disconfirming ones: I anticipate that the reddish color I now see will be continued on the other side; I anticipate that the rough and natural-looking shape

Determinable Indeterminacy  131 it now shows me will be continued as well. Only on the basis of such anticipations can my subsequent experience of the thing be of that very thing. Had I no such anticipation, the color and shape of the back side (whatever they might be) would simply be the content of a new experience.20 But with the anticipation, there is a meaning that prescribes the determinability of the thing’s indeterminateness in such a way that I can increase my knowledge of it. As I move around the formation, I confirm its reddish color, and this subsequent experience has the character of getting to know the thing as it already was. Similarly, if the color turns out to surprise me, that color belongs within the range of anticipated possibility, and though my specific anticipation is thwarted, I still learn how the rock formation was already determined. Phenomenologically, such experiences are not explorations of the thing; they are explorations of the horizons of the thing. Looking to the external horizon, my perception of the rock formation is framed by the co-presence of rock walls and sand, and I might anticipate that the formation’s shape has arisen through ancient water-flows and erosion. But if I move around it and find that, while the reddish color is confirmed, the formation is not rough and natural-shaped but rather smooth and planar with what look like inscriptions on it, my partial intention of the shape—and with it, perhaps, what the thing itself is—has been disconfirmed and another has taken its place. This change in the internal horizon brings with it a different set of intentional implications that guide me in regard to the more or less indeterminate external horizon: This thing belongs to the world not as something merely naturally formed, but as something that has a meaning deriving from practices of which I know nothing. Even in my ignorance, however, that meaning, and those practices, are themselves anticipated, according to Husserl, in certain typical ways: I am familiar with social practices (though perhaps not the ones responsible for the shape and markings) and this sort of indeterminacy is in turn determinable through various measures—some perceptual, others much more complicated—that I can take to learn more. For Husserl, then, familiarity is phenomenologically basic. This means that “the difference between the familiar and the new is relative”; indeed, “an absolutely unknown thing is something contradictory.”21 In circumnavigating the rock formation, I might find that it disappears altogether. It would thereby be struck from the world of actuality—it is not what it presented itself as being—but not from the horizon of experience altogether. A new external horizon would open up, and I would perhaps search for an explanation: it is a cleverly crafted hologram; it was an illusion occasioned by the angle of the sun’s rays; I ingested an hallucinogen but forgot about it. In all such cases, I cannot experience the thing without a prior type against which the confirmatory or disconfirmatory character of subsequent experiences can be measured. On the other hand, “absolute determinateness is [. . .] only an ideal limiting case.”22 I never come to the end of exploring the internal and external horizons of a thing.

132  Steven G. Crowell Here we see that because indeterminacy is always bound, perception is meaningful. The internal horizon of a perceptual object is bound, at a minimum, by the “generic form” of the object: rock formation, natural kind, or perhaps just “spatial object.”23 But the external horizon, while limitless, is also bound: the “ ‘form’ of the world”—its “unalterable identity” as actual—prescribes the ontological coherence of my experience. This generic type, and this world-form, as features of the perceptual horizon, are not themselves things but rather normative points of reference that allow things to serve as measures of what we say or see of them. To perceive something “as” a rock formation is possible only within a horizon, but this is fortunate because only so can the way the thing shows itself from over there serve the epistemic function of giving me a “better view,” of being a further determination of what was initially determinably indeterminate. For phenomenology, determinacy and indeterminacy belong to the horizon, and the horizon is what it is only because of its teleological (normatively regulative) form as a space of meaning. But what about this sort of meaning? Isn’t it just a finite version of what, at the limit, would be determinateness independent of any reference to “subjective” horizons? If it were, there would be no conflict between Husserl’s claim that what belongs to the horizon is never fully determinable and his claim that the form of the world guarantees the actuality—the metaphysical determinateness—of everything in the world. He could simply argue that while our “conceptions” of things as a whole are always on the way, the concept of worldly actuality entails metaphysical determinateness—that is, a sort of determinateness that could be asserted without reference to any horizon. But what could authorize this move from a phenomenology of horizons to a metaphysical conclusion? In the course of many decades devoted to thinking about this question, Husserl never, it seems to me, came to a fully satisfactory answer. The reduction yields a metaphysically neutral standpoint that sticks to the immanence of experience’s “infinite task,” but in his more rationalist moments Husserl seems to opt for conceptualism: the world’s “unalterable identity” as actual seems to entail the metaphysical determinateness of everything in the world. Reason catches up with the real in the infinite, so to speak, but never in experience. Pursuing the details of Husserl’s metaphysics, however, would take us far beyond the scope of our present concerns.24 Regarding the phenomenology of determinable indeterminacy itself, we must acknowledge that Husserl’s account of the meaningful character of the horizon—an account that appeals to bodily engagements with the world, such as hiking around a rock formation to get a better look—complicates the picture. What does the phenomenology of such “engagement” contribute to constituting the horizon as the space of meaning in which something can be experienced as determinably indeterminate?

Determinable Indeterminacy  133

Merleau-Ponty on the Horizon: Ontological Ambiguity Merleau-Ponty begins Part I of his Phenomenology of Perception by reflecting on an apparent implication of Husserl’s concept of horizon. If perception “terminates in objects,” and if the object thus “appears as the reason” for all the experiences we have of it, then two metaphysical options for characterizing the object seem open to us. One option holds that truth (determinateness) is the object “seen from nowhere,” i.e., its being the reason for our perception of it is only contingently connected to perception.25 The other option holds that the object is the ideal unity of all the veridical experiences— perception, memory, valuation—that I or anyone might have of it. The first option, objectivism, eliminates any epistemic role for perception; the second, phenomenalism, replaces the object with a subjective synthesis. The phenomenology of horizons, on Merleau-Ponty’s view, undercuts this metaphysical dilemma. The phenomenalist forgets that my experience of the unity of the object is prior to (and so cannot be accounted for by) a synthesis: appeals to “express memory” and “explicit conjecture” could only yield “a probable synthesis, whereas my perception [i.e., the perceived] is given as actual.”26 When I attend to something in my perceptual field, the things around it “recede into the margins and become dormant, but they do not cease to be there,” and “the object I am currently focusing on—seen peripherally—is implied in” the “other horizons” that belong to what has so receded. Thus, it is the horizon itself, and not any subjective synthesis, that “assures the identity of the object throughout the exploration.”27 The objectivist’s view from nowhere, in turn, fails to recognize that the determinations of the object in question are perceptual ones, and thus horizonal. As Merleau-Ponty playfully suggests, a static perception experiences the object’s unity, determinateness, as what is already there for the gaze of “other things.” If part of the object is occluded, this does not mean that it is not “revealed” in itself, since it already includes the properties that “the fireplace, the walls, and the table [that surround it] can ‘see.’ ” The horizon thus assures the identity of the object through my subsequent exploration because the object itself is “what is seen from everywhere”—i.e., nowhere. The house “has its water pipes,” even though “we may never see them.”28 If the determinateness of the object can be understood neither as a subjective synthesis nor as a view from nowhere, we must make a new start, phenomenologically, with the object in a horizon, the object viewed from somewhere. For Merleau-Ponty, reflection on the horizon—in particular, on the dynamical character of perception as exploration of the horizon—demands that we take seriously an insight implied in Husserl’s description but not highlighted: in exploring the horizon we are also in it, by means of the motor capacities of our bodies. There is neither a pure intellectual grasp of the object itself nor a view from everywhere; rather, the focal object always

134  Steven G. Crowell appears as a Gestalt, a figure against a ground or “halo” of indeterminateness, as Husserl noted. But this horizon is not a kind of cinematic screen that merely faces us;29 it is a Gestalt formed by the experienced embeddedness of my body in the world, by the way the former “inhabits” the latter. The visual notion of figure-ground is an abstraction from what MerleauPonty calls the “body schema”—not “an experience of my body” but of “my body in the world”30—thanks to which I inhabit the world through a “system of equivalences,” a tropological, figurative, or analogical mode of transposition that renders unfamiliar situations familiar on the basis of my acquired abilities. If, for Husserl, the perceptual horizon is structured by intentional implications that afford the normative anticipations which enable further determination of the perceived, then for Merleau-Ponty this normativity is inexplicable unless such anticipations derive from the body’s “motricity” (the “I can”), its exercise within a horizon to which it, too, belongs. Such a Spielraum is not “objective space” but “world” as “my body’s inseparable correlate.”31 Motricity makes the correlate a correlate: “Bodily space and external space form a practical system, the former being the background against which the object can stand out.”32 It also provides the horizon with its normative character. Quoting Grünbaum, MerleauPonty writes: “motricity already possesses the elementary power of sensegiving (Sinngebung);” it is the “primary sphere” in which the “sense of all significations” is first given.33 Motricity, then, the “I can,” is “original intentionality.”34 Merleau-Ponty thus preserves the core of Husserl’s notion of horizon, its character as a space of meaning connected by intentional implications that are neither causal nor intellectual. However, his conception of motor intentionality transforms the relation between the horizon (as the never fully determinable indeterminacy of what we encounter in the world) and the world’s “form,” its unalterable identity as actual. This transformation appears most clearly when we consider how it is that motor intentionality is able to open up a horizon in the first place. Though it inhabits the world, the body is not simply an object in it; rather, it is a schema of abilities that requires us to acknowledge a “new type of existence” which eludes the traditional categories of body and mind.35 My body’s spatiality is not “positional” but situational; to be “here” is not to occupy an objective point but to “install” a matrix of coordinates whose values are determined by what I can do. The body schema is the way “the parts of the body” are integrated into a unity “according to their value for the organism’s projects,” an integration whereby the world becomes the correlate of those projects, the ground against which things emerge with a kind of significance that is “polarized by [the body’s] tasks.”36 Motor intentionality discloses significance as what affords such projects; the world is inhabited as a space of “possibilities,” a “solicitation” of bodily adjustments “which is exerted upon the body without any representation.”37 Affordances do not cause, but motivate, such adjustments and so can be

Determinable Indeterminacy  135 normatively assessed in terms of the organism’s success or failure.38 The door handle affords me entrance to the room without any “partial intentions” that take it as a thing having such a property; the dark corner may afford me a hiding place, though, as Sartre points out, it also affords my pursuer an obvious place to look.39 This last point—let us call it the ambiguity of affordances—is crucial to Merleau-Ponty’s transformation of Husserl’s phenomenology of the horizon. If the normative order of the horizon—that which makes things within it meaningful, determinably indeterminate—exerts its solicitations without any representation, then it cannot be analyzed in terms of partial intentions aiming at what future perceptions must be like if the thing is as I take it to be. What sort of normativity, then, characterizes the Gestalt formed by the body’s motor-intentional projects? Here Hubert Dreyfus has introduced an important distinction between conditions of satisfaction and conditions of improvement.40 Because motor intentionality precedes all representation, all mental content, Dreyfus argues that the normativity informing “comportments”—the skillful coping or bodily habits exercised without deliberation and without effort (trying), as in the “flow” of a tennis match or in my routinized drive to work—cannot be understood in terms of logico-semantic conditions of satisfaction. Absorbed in a tennis volley, I am not guided by partial intentions that specify where best to move in order to return the ball; rather, “the phenomenal forces at work in my visual field obtain from me, without any calculation, the motor reactions that will establish between those forces the optimum equilibrium.”41 As Dreyfus explains it, these reactions are attuned to normative conditions of improvement. The skilled body “knows” how to improve its situation in relation to these horizonal features; its project follows a path of “felt approximation” to an optimal body/world Gestalt.42 Optimality, while normative, is thus inseparable from the particular situation; it cannot be generalized, and yet it exercises a meaning-constituting function and requires that we recognize, as Merleau-Ponty puts it, a “new sense of the word ‘sense.’ ”43 For our purposes, what is most important about this new sense of “sense” is that it belongs to the phenomenological field, an “ambiguous domain.”44 The determinable indeterminacy of the horizon does not rest upon “world” as the unalterably identical “form” of actuality but on an irreducibly ambiguous mode of being: “world” as the correlate of an ever-shifting Gestalt keyed to the “intentional arc” of my motor projects.45 Here Merleau-Ponty resolves Husserl’s aporia of indeterminate horizon and determinate world in favor of the indeterminacy of the horizon. While he nominally agrees with Husserl that “everything resides within the world,”46 things in the world partake of the world’s own ontological ambiguity. Further consideration of Dreyfus’s analysis reveals how deeply this ambiguity goes. In order to describe the way comportment is attuned to its conditions of improvement, Dreyfus borrows a term from Kant’s reflections on

136  Steven G. Crowell the work of art: comportment is “purposive action without a [represented] purpose.”47 This is how Merleau-Ponty understands his own phenomenological project as well: Philosophy is not “reflection of a prior truth” but, like art, “the actualization of a truth.”48 If bodily intentionality is the “existential core of signification,” it is nevertheless not governed in advance by reason, be it third-person machine-logic or first-personal representation. Rather, the “genesis of meaning” lies in the “assimilation of accidents”— worldly encounters—“in order to construct a reason from them.”49 Existence itself—“the perpetual taking up of fact and chance by a reason that neither exists in advance of this taking up, nor without it”—belongs to this realm of ambiguity.50 Bodily existence may be “condemned to sense,” but this sense is nowise governed by a logos. Comportment is responsible for the manner in which “perceptions confirm one another,” but the phenomenological world in which this takes place is not “pure being,” an unalterable form that guarantees actuality. Rationality is established “through an initiative that has no ontological guarantee.”51 Comportment thus opens up horizons of significance in much the way an artist reveals a world through the “accidents” of her brushstrokes. The conditions of improvement which normatively guide such practices partake of the same ambiguity that characterizes the new type of existence discovered in the body schema itself. But if in this way Merleau-Ponty’s phenomenology of the horizon seeks to avoid the twin metaphysical dead-ends of objectivism and phenomenalism, it might appear that the ambiguity of being undermines the possibility of determinateness altogether. In Merleau-Ponty’s view, the phenomenology of horizons shows that “the only logos that pre-exists is the world itself.”52 But if such pre-existence does not possess the unalterable “form” of actuality—if the necessary indeterminacy of the horizon is understood as the irrevocable ambiguity of being—then inquiry (science, philosophy), the conceptual determination of what is, threatens to become incomprehensible. If Husserl tended toward rationalism in addressing the issue of how the essential indeterminacy of the horizon relates to the world’s unalterable form of actuality, then Merleau-Ponty seems to turn in the other direction: the essential indeterminacy of the horizon is ontologically basic; the body is an emblem of ontological ambiguity. Inquiry not only cannot come to an end (in the sense of Husserl’s “infinite task”); it belongs to a world where the determinateness of things is ruled out. Considering the outcome of Merleau-Ponty’s transformation of Husserl’s horizon-concept, Emmanuel Levinas wondered whether “the objectivity of the object is [not] thereby underestimated”: what appears, on Husserl’s view, as the determinable indeterminacy of things becomes, in Merleau-Ponty, fundamentally equivocal, an ersatz determinacy correlated to a vast plurality of “worlds” or milieux of embodied organisms.53 To highlight the problem here we may return, once more, to Dreyfus. Of the “new sense of the word ‘sense’ ” which comportment, as motor intentionality, requires us to acknowledge, Dreyfus notes that it “covers both

Determinable Indeterminacy  137 animals and people,” and he contrasts this with the “special kind of significance” available to human beings: “We might call this meaning.”54 What does Merleau-Ponty tell us about such meaning? And can this distinction explain how a non-equivocal determinateness of things is compatible with a phenomenological ontology of ambiguity? Here Merleau-Ponty introduces his concept of “expression.”55 Because the body is always out for something in its projects, it is sensitive to the affordances correlated with its striving, and so bodily movements are “gestures” that “signify” how the whole inhabited Gestalt is structured. All organisms are “expressive” in this sense. But while the animal organism’s projects are instinctual, the activities of the human being “break free from animal life,”56 and so human gestural expressions take on a “form of generality”—i.e., of iterable Gestalts sensitive to various sorts of conditions of improvement based on habitus or skill acquisition.57 Sometimes the body “restricts itself to gestures necessary for the conservation of life, and correlatively it posits a biological world around us.”58 At other times, “playing upon these first gestures and passing from their literal to their figurative sense, it brings forth a new core of signification through them”—for instance, in dance. But when “the signification aimed at cannot be reached by the natural reach of the body,” we “construct an instrument, and the body projects a cultural world around itself.”59 It is in this cultural world that Dreyfus’s distinction between animal significance and human meaning takes on salience. In particular, one form that the body’s gestural expression can take is speech, a bodily capacity that yields words the way the organism’s motricity yields affordances.60 Speech alone, of all bodily capacities, constitutes an “intersubjective acquisition;” in speaking and writing, one is “aware of intending the same world with which other [speakers and writers] were already concerned” and so it “instills in us the idea of truth” as the “presumptive limit of its efforts.”61 Such an idea is, presumably, that of a determinate reality that would not be measured by the organism’s striving to achieve an optimal Gestalt in an environment but by the world itself as it is in itself. But two questions go unanswered here: First, how does language provide us with such an idea of truth? And second, how can the limit idea of a determinate reality be squared with the ontological ambiguity of the phenomenological domain? Though he struggled with these questions right up to his last, unfinished work, The Visible and the Invisible, Merleau-Ponty never found an adequate answer. The various significance-producing environmental encounters granted to different orders of animal organisms, including human beings, yield a plurality of logoi, a plurality of incommensurable “worlds” or niches, and Merleau-Ponty’s claim that “everything resides within the world” remains equivocal. The kind of indeterminacy characterizing such “worlds” does allow for significance, or affordances, but it fails to surmount the ambiguity of affordances. Indeed, “the absolute positing of a single object is the death of consciousness.”62 Reason can get no grip

138  Steven G. Crowell on the determinateness of things because concepts are at bottom expressive gestures: “speech is a gesture, and its signification is a world”—but not the world.63 The phenomenology of objectivity has indeed been “underestimated” here. This, as I will argue below, is because projects have not been given their due. If we want to understand the determinable indeterminacy of the horizon, we must start with what Dreyfus called human “meaning” and not with the kind of bodily intentionality we share with animals. Dreyfus is right that “originary intentionality” belongs to comportment, but the latter is not exhausted by skillful coping.64 This point leads us to Heidegger, whose phenomenology of being-in-the-world distinguishes sharply, as MerleauPonty does not, between animal “striving” and human “care” for the meaning of its own being. He is thus able to supplant Merleau-Ponty’s ontology of ambiguity by means of a two-tiered account of comportment in which are preserved both the possibility of truth or objectivity (determinateness) and the necessary indeterminateness that belongs to the disclosure of the horizon as a meaningful one.

Heidegger: Horizonal Normativity and Ontological Indeterminacy Heidegger describes the aim of Being and Time as the explication of “time as the transcendental horizon for the question of being.”65 In contrast to Husserl and Merleau-Ponty, then, Heidegger does not introduce the horizon as an element of perceptual experience—the halo of indeterminacy surrounding each perceived object, the perceptual Gestalt formed by motor intentionality—but as a condition of Dasein’s understanding of being. The text culminates in the claim that our various ontological concepts—reality, ideality, subsistence, and so on—are grounded in the “ekstatic-horizonal temporality” of Dasein’s care-structure, a “schema” that informs the intelligibility of things.66 But this does not mean that the phenomena considered by Husserl and Merleau-Ponty find no place in Heidegger’s analysis, since he also occasionally uses the term “horizon” in connection with his own concept of world. As we saw, Husserl introduced the concept of horizon in the context of a description of the natural attitude whose point was to motivate a radical change of attitude—the phenomenological epoché—necessary for the philosophical elucidation of how meaning in the natural attitude is constituted for consciousness. Heidegger begins in precisely the same way, with a description of what he calls Dasein’s “average everydayness.” The aim is to determine what “world” is in the “horizon of average everydayness.”67 World is the “totality of significations” (or intentional implications) thanks to which things can show up in the world determinately, “as they are in themselves.”68 And, like Husserl, Heidegger sees this sort of intelligibility— “the ‘natural’ horizon for starting the existential analytic of Dasein”—not

Determinable Indeterminacy  139 as sui generis but rather as “only seemingly self-evident.”69 Phenomenology must subject it to a reduction in order to grasp the ontological (transcendental) conditions that make it possible.70 And because Heidegger, like MerleauPonty, sees that the being who has such a world-horizon must itself belong to it, he begins his analysis of those conditions with an account of Dasein’s practical engagement with things in the world, its “circumspective concern.” Unlike Merleau-Ponty, however, Heidegger’s emphasis is not on the bodily Gestalt this entails but on the normative order of the world-horizon, which makes determinate intentionality—taking something as something— possible. In the two-tiered conception of agency this involves, as we shall see below, ambiguity is no longer a feature of the world itself but a necessary moment in the disclosure of meaningful horizons, “worlds,” through Dasein’s projects. Before getting to that, we should note that Heidegger’s approach does not resolve Husserl’s aporia: it remains neutral with respect to the question of whether the determinable indeterminacy of the horizon (world) entails metaphysical determinateness. The issue was especially important to Husserl, since for him the primary significance of the world-horizon was its unitary character, its being “spread out endlessly in space and time,” containing everything that in any way is. Heidegger is not unaware of this conception of world—which he terms das Seiende im Ganzen—and it is even alluded to in Being and Time.71 But for him it is a metaphysical concept that can be elucidated only on the ground of the transcendental phenomenology of Being and Time.72 Our concern, then, will be exclusively with that text, and with Heidegger’s account of how Dasein’s comportment discloses the world-horizon as a normative space of meaning that makes the determination of something as something possible. Heidegger’s analysis of “comportment” (Verhalten)—his term for the “primordial intentionality” that belongs to Dasein’s average everydayness— includes the sort of skillful coping or motor intentionality that belongs to the unreflected “flow” of my bodily engagement with things. The skilled carpenter effortlessly wields the tools and materials that afford various moves in the project of building a birdhouse, and such effortlessness is guided by a knack for finding the optimal position for doing a certain sub-task, a sensitivity to conditions of improvement that will establish an equilibrium between body and world. This aspect of comportment, then, excludes the effort of “trying” (representing each step and assessing what is optimal in relation to it), as Dreyfus remarks, and also takes place without any rational deliberation, which arises only when something goes wrong. In another sense, however, comportment is a matter of trying: I am trying to do something—namely, make a birdhouse. Without this, the things that afford success as I effortlessly manipulate tools and materials remain ambiguous: If what I am trying to do is make a birdhouse, then a very large hammer and a set of ten-penny nails will not afford or enable the task, whereas if, in trying to make an artwork, I engage in the very same

140  Steven G. Crowell movements, with the very same tools and materials, the affordances will be sensed differently. In Heidegger’s language, my bodily manipulation of tools and materials always takes place in an “equipmental totality,” a horizon of intentional implications—Heidegger’s term is Verweisungen—that refer tools (hammers) to other tools (nails) and to various sorts of material etc.73 And this equipmental totality is itself determined by the “work” to be done, by what I am trying to do with them.74 It is the work that determines what tools and materials are appropriate for the job, what affords what, and so my trying to do some specific thing is what disambiguates the ambiguity of affordances. In Heidegger’s analysis of comportment, “trying to do” represents only one dimension or tier, however. The need for a second tier becomes clear as soon as we ask what it is that makes it the case that I am trying to make a birdhouse rather than an artwork. It cannot be the mere “representation” of an outcome, since that outcome is itself ambiguous: is this thing I produced a successful artwork or a failed birdhouse? On Heidegger’s view, the success or failure of what I am trying to do—and so the equipmental totality in which things show up as suitable or appropriate for doing it—is determined by an altogether different dimension of comportment: my trying to be something. My skillful coping with things yields determinate affordances only if I am at the same time exercising a different sort of skill, “competence over [. . .] being as existing.”75 If I am trying to be a carpenter—that is, if I am sensitive to the norms of carpentry and care about living up to them— then what I make will be judged as successful or unsuccessful in terms of birdhouses, and not in terms of what would be appropriate were I trying to be an artist. In Heidegger’s language, the “in order to” structure of the equipmental horizon remains ambiguous until we can specify that “for the sake of which” I am doing what I do, and the “for the sake of which” has “no further involvement” (it is not part of the equipmental totality); rather, it pertains to “a possibility of Dasein’s being,” where “possibility” means an ability to be something: carpenter, artist, teacher, parent, and so on.76 For Heidegger, then, comportment consists in the unity of trying to do and trying to be. This is what he calls a “project.” In animals, as MerleauPonty suggested, the role of trying to be is taken over by “instinct,”77 but Dasein is distinguished from animals by its “understanding of being,” that is, by the fact that “in its very being that being is an issue for it.”78 To say that its being is “an issue” does not mean that Dasein now and then worries about who it is; rather, to exist (in Heidegger’s technical sense) is to take a stand on what it means to be whatever it is that Dasein is trying to be. This stand-taking—“commitment” as binding oneself to a normative exemplar (Vorbild) of a way of being—opens up the horizon or discloses “world” in Heidegger’s sense. The world of carpentry or teaching or parenthood is disclosed only for one who is able to act for the sake of being a carpenter, teacher, or parent. As the “totality of significations” relevant to teaching or parenting, the world is sustained by my care or concern for what it means to

Determinable Indeterminacy  141 be a teacher or parent, i.e., my commitment to acting in light of the norms that determine such a way of being. Within that world, things can take on definite or determinate significance. But what sort of norms are those, and how do they determine success or failure? Here we find the most important consequence of Heidegger’s two-tiered analysis of comportment, since the norms of trying to be are significantly different from the norms of trying to do. If my commitment has established what I am trying to be, then the process of trying to do can be understood as something like a recipe: given the task of making a birdhouse, there are (more or less) determinate steps for reaching a goal whose success is also (more or less) determinate. This is because the “work” to be produced—be it a birdhouse, a victorious tennis match, or successful navigation of the drive home—provides a measure of success or failure that lies “outside” the activity of producing it. But this is not the case for trying to be something, whose measure is always at issue. Trying to be something is atelic—that is, while it has an “end in view,” that end is not something to be reached in the future but rather is “at its end” at every moment of the trying. Thus it is continually succeeding or failing, while the measure of such success or failure is itself precisely what is at stake in my understanding of what it means to be a teacher, parent, etc. It might seem that there is no real distinction between the norms of trying to do something and those of trying to be something. If there are recipes for making houses, are there not recipes for being a teacher? In a sense, there are; but to see why this does not compromise the distinction we must reflect on what Merleau-Ponty called the “world of culture” and Heidegger calls das Man. The world-horizon for Heidegger is possible only because Dasein is always “with others”—not in the sense that it finds itself among conspecifics but in the sense that it understands itself within a social world where there are standard uses for things and normal ways of behaving.79 Comportment always inhabits a world-horizon where the “typical” and “average” have already determined a kind of intelligibility for things. For this reason, too, the world-horizon is not a matter of affordances (significance) but permits an explicit as-structure (meaning): things have conditions of satisfaction and not merely conditions of improvement. The hammer does not just afford hammering; it is understood to be something used in order to do something. It is available as a hammer, “articulated” as such, even if I make no use of it. The hammer is determinately a hammer neither because it is called one around here, nor because it affords hammering. A rock affords hammering, but it is neither called a hammer nor is it one. I can use a rock as a hammer, but that doesn’t make it one. Rather, something is rightly called a hammer because its being lies entirely in its appropriateness for certain uses within a context of other things whose being is similarly determined by such appropriateness. Now, among these “things” are also common understandings of what teachers or parents should be, how they should act, what they normally

142  Steven G. Crowell do. Being a teacher or a parent is, in this sense, a social-normative status, and the institutional and social constraints on what constitutes teaching or parenting can be treated as recipes for success. But this is precisely the sort of merely apparent self-evidence that phenomenology, in Heidegger’s view, must see through. Understood phenomenologically, parenting and teaching are “possibilities for being a self” whose meaning—how best to go on—is always at issue. Thus the social-normative status of being a teacher is both a necessary starting point and a trap: if I treat the norms and expectations involved as recipes—act as “one does” simply because it is what one does—I abdicate my responsibility for deciding whether one ought to go on in that way. In being “transparent” (authentic) with regard to this responsibility, I might well find that I continue to do what one does, but I may also find that I must part ways with consensus, perhaps radically. The phenomenological point is not whether I do or do not conform; rather, it is that only such being responsible (which Heidegger calls “taking over being a ground” or “resoluteness”) can ground the meaning that infuses the world-horizon and makes things within it determinably indeterminate.80 The norms at issue in being a teacher are what make it the case that the same lesson plan which is an “exemplary instance of the genre” under one understanding of what it means to be a teacher will be a “hindrance to learning” under another. Both will be entirely determinate in their respective worlds, but the worlds themselves—the horizon disclosed in an understanding of being—will not be determinate in the same way. For Heidegger, then, “world” is always the world of an understanding of what it means to be something (teacher, parent), a horizon of intentional implications made possible by my orientation toward what is best in the matter of teaching or parenting. Thus, Dreyfus is right to describe comportment as “purposive action without a [represented] purpose,” though such purposiveness is not exhausted by bodily intentionality; instead, as yielding (human) meaning, it includes acting for the sake of being something. My going on in a particular way has the character of what Kant called “exemplary necessity,” i.e., it is judged as an example of a “universal rule that we are unable to state.”81 Heidegger makes the same point in Platonic language: acting for the sake of something—“sovereignty over itself as hou heneka”—is the “essence of the agathon,” which in turn is what makes “truth, understanding, and even being” possible. But because this normativity (what is best in the matter of what I am trying to be) is “beyond being” (epekeina tes ousias), it is “not by accident that the agathon is indeterminate with respect to its content, so that all definitions and interpretations in this respect must fail.”82 The determinable indeterminacy of what I am trying to do—the horizon of the work—rests upon an indeterminable indeterminacy—not of the world but of the norm (ta agathon) at issue in my decision for a certain way of best going on in the

Determinable Indeterminacy  143 matter of what I am trying to be. This is what distinguishes human meaning, in Dreyfus’s sense, from animal significance. If Heidegger thus displaces Merleau-Ponty’s notion of ambiguity from being itself into the structure of comportment, the “understanding” of being, we might well wonder why such a displacement does not also compromise the very idea of the determinable indeterminacy of the world-horizon. Has objectivity really been given its due? We cannot provide a full answer here. That would require investigating Heidegger’s phenomenology of scientific inquiry, in which the practical world-horizon of circumspection is replaced by the artificial meaning-structure of “theory” (for instance, a “mathematical projection of nature”). This, in turn, allows for a practice which “thematizes” entities, i.e., frees them “in such a way that they can ‘throw themselves against’ a pure discovering—that is, that they can become ‘objects.’ ”83 We will wrap up this note on the phenomenology of horizons simply by pointing out that a theory is itself a kind of equipment whose appropriateness derives from my scientific comportment, my trying to be a scientist. As Heidegger describes it, the commitment belonging to such comportment involves at least one meaning of “objectivity”: science is “exceptional” in that “it gives the matter itself explicitly and solely the first and last word.”84 At the same time, how best to go on in the matter of science always remains at issue.85

Notes 1 Husserl (1989, p. 408). 2 Husserl (1983, p. 56). 3 Ibid., p. 51. 4 Ibid., p. 51. 5 Ibid., p. 51. 6 Ibid., p. 52. 7 See Natanson (1968, p. 55), where Natanson refers to a passage in James Joyce’s A Portrait of the Artist as a Young Man. 8 Husserl (1983, p. 52). 9 On “intentional implications” see, for instance, Husserl (1959, pp. 144f, 149). Husserl also calls them “Geltungsimplikation der Horizont” (p. 150), emphasizing their normative character. 10 Husserl (1983, p. 53). 11 Ibid., p. 55. 12 Ibid., p. 57. 13 For some discussion of the literature, and my own view, see Crowell (2013, chs 2–3). 14 Husserl (2004, p. 62). 15 Ibid., p. 59. 16 On internal and external horizons, see Husserl (1970, p. 162). For a more detailed elaboration of the determination of perceived things via their internal and external horizons—which Husserl calls “explication”—see Husserl (1973, pp. 112–129). 17 Husserl (2004, p. 60). 18 On the noema, see Drummond (1990).

144  Steven G. Crowell 19 As Husserl already writes in 1911: “All types of consciousness” are, “so to speak, teleologically ordered under the title of knowledge,” and phenomenology must explore their “essential connection and their relation back to the forms of the consciousness of givenness belonging to them” Husserl (1965, p. 90). 20 As Husserl writes in Philosophy as Rigorous Science (1965, p. 87), the question to which phenomenology provides the answer is this: “How can experiences be mutually legitimated or corrected by means of each other, and not merely replace each other?” 21 Husserl (2004, pp. 61–62). 22 Ibid., p. 61. 23 Ibid., p. 62. 24 On Husserl’s metaphysics, see Tengelyi (2014), De Santis (2018), Smith (2003), Loidolt (2015, pp. 103–135), De Palma (2017, pp. 1–18), and Zahavi (2018). 25 Merleau-Ponty (2012, p. 69). 26 Ibid., p. 70. 27 Ibid., p. 70. 28 Ibid., pp. 71–72. Later, and less playfully, Merleau-Ponty will argue that this all-sidedness is a function of intersubjectivity: the determinateness of the unseen side is guaranteed by the presence of others; it is what the other sees, or could see, while I remain in place. But this hardly solves the problem, since the other cannot be “everywhere” either. 29 Ibid., p. 70. 30 Ibid., p. 142. 31 Ibid., p. 143. 32 Ibid., p. 105. Maxine Sheets-Johnstone, The Primacy of Movement (New York: John Benjamins, 2011), develops this point in detail and also argues that Husserl’s own analyses of embodiment are, in many respects, more detailed and perspicuous than Merleau-Ponty’s. 33 Merleau-Ponty (2012, p. 143). 34 Ibid., p. 139. 35 Ibid., p. 102. 36 Ibid., pp. 102–103. 37 Ibid., p. 140. The notion of “affordances” was introduced by the psychologist J. J. Gibson, whose approach to perception was influenced by MerleauPonty. Affordances are situation-specific ways in which the world provides us with openings for action that are not objective features of the world but features of the “complementarity of the animal and the environment.” See Gibson (1979, p. 127). Dreyfus has made affordances central to his characterization of the indeterminacy of the horizon. In contrast to “conceptualists” like John McDowell, who see the world as “the totality of objects, events, and states of affairs,” Dreyfus argues that the world is first of all a space of affordances: “we directly perceive affordances and respond to them without beliefs and justifications being involved. Moreover, these affordances are interrelated, and it is our familiarity with the whole context of affordances that gives us our ability to orient ourselves and find our way about.” See “Overcoming the Myth of the Mental” in Dreyfus (2014, p. 120). 38 On motivation see Wrathall (2005). 39 Sartre (1956, p. 353). 40 Dreyfus (2014, p. 150) in the essay “The Primacy of Phenomenology Over Logical Analysis.” 41 Merleau-Ponty (2012, p. 109). 42 The term “felt approximation” In Husserl’s: See Husserl (2004, p. 145). 43 Merleau-Ponty (2012, p. 148).

Determinable Indeterminacy  145 4 Ibid., p. 65. 4 45 Ibid., p. 137. 46 Ibid., p. 204. 47 Dreyfus (2014, p. 88) in the essay “Heidegger’s Critique of the Husserl/Searle Account of Intentionality.” 48 Merleau-Ponty (2012, p. lxxxiv). 49 Ibid., p. lxxxiii. 50 Ibid., p. 129. 51 Ibid., p. lxxxiv. 52 Ibid., p. lxxxiv. 53 Levinas (1969, p. 94). On the kind of “equivocation” at issue, see the discussion on pp. 90–98. 54 Dreyfus (2014, p. 259) in the essay “Why Heideggerian AI Failed and How Fixing It Would Require Making It More Heideggerian.” 55 Merleau-Ponty (2012, p. 142). 56 Ibid., p. 195. 57 Ibid., p. 147. 58 Ibid., p. 147. “Already the presence of a living being transforms the physical world, makes ‘food’ appear over here and a ‘hiding place’ over there” (p. 195). 59 Ibid., p. 148. 60 Ibid., p. 189. 61 Ibid., p. 196; my emphasis. 62 Ibid., p. 74. 63 Ibid., p. 190. 64 Dreyfus is not unaware of this, of course, but he consistently assimilates comportment to skillful coping. In “Heidegger’s Critique of the Husserl/Searle Account of Intentionality” he writes: “The phenomenon of purposive action without a [represented] purpose is not limited to bodily activity but pervades all types of comportment.” See Dreyfus (2014, p. 86). But he also identifies the key aspect of comportment from which meaning arises—our “understanding of being”—with a “general background coping, our familiarity with the world” (pp. 90–91). However, as we shall see, the normative order that pertains to our understanding of being cannot be adequately understood either in terms of coping or in terms of familiarity. 65 Heidegger (1962, p. 65). 66 This Kantian aspect of Heidegger’s project was never brought to completion, and there are serious difficulties with it. For an excellent account, see Golob (2014). 67 Heidegger (1962, p. 94). 68 Ibid., p. 120 69 Ibid., p. 423. 70 Heidegger’s reduction is not identical in every respect to Husserl’s of course. See his comments on this matter in Heidegger (1982, p. 21), and the discussion in Crowell (2013, pp. 72–77). 71 See, for instance, Heidegger (1962, p. 292). 72 Heidegger pursues this metaphysics in writings and lectures between 1928 and 1935, and he encounters many of the same problems that beset Merleau-Ponty’s ontology of ambiguity. But he abandons the project—for good reasons, in my view—and exploring the question of whether he resolves Husserl’s aporia in these writings would take us too far afield here. For some discussion, see Crowell (forthcoming). 73 Heidegger (1962, p. 107). 74 Ibid., p. 99: “The work bears with it that referential totality within which the equipment is encountered.” 75 Ibid., p. 183.

146  Steven G. Crowell 76 Ibid., pp. 116–117. For more detailed discussion of these points see Crowell (2013, Part III). 77 Merleau-Ponty (2012, p. 195). 78 Heidegger (1962, p. 32). 79 Ibid., pp. 156, 164–167. 80 Ibid., pp. 330, 343. For more on responsiveness to norms, transparency, and authenticity, see Crowell (2013, chs. 8–9). 81 Kant (1987, p. 85). 82 Heidegger (1998, p. 124) in “On the Essence of Ground.” 83 Heidegger (1962, ¶69b; p. 414). 84 Heidegger (1998, p. 83) in “What Is Metaphysics?” 85 For an attempt to spell out Heidegger’s notion of commitment in the context of the project of science, see Haugeland (2007) and (2000). For a sympathetic critique of Haugeland and an alternative approach to what is at stake in science, see Rouse (2002) and (2015).

Works Cited Crowell, Steven. 2013. Normativity and Phenomenology in Husserl and Heidegger. Cambridge: Cambridge University Press. ———. (forthcoming). “The Middle Heidegger’s Phenomenological Metaphysics,” in The Oxford Handbook of the History of Phenomenology, edited by Dan Zahavi. Oxford: Oxford University Press. De Palma, Vittorio. 2017. “Phänomenologie und Realismus. Die Frage nach der Wirklichkeit im Streit zwischen Husserl und Ingarden.” Husserl Studies 33(1): 1–18. De Santis, Daniele. 2018. “ ‘Metaphysische Ergebnisse’: Phenomenology and Metaphysics in Husserl’s Cartesian Meditations §60. Attempt at Commentary.” Husserl Studies 34(1): 63–83. Dreyfus, Hubert. 2014. Skillful Coping: Essays on the Phenomenology of Everyday Perception and Action, edited by Mark Wrathall. Oxford: Oxford University Press. Drummond, John. 1990. Husserlian Intentionality and Non-Foundational Realism. Dordrecht: Kluwer. Gibson, James J. 1979. The Ecological Approach to Visual Perception. Boston, MA: Houghton Mifflin Harcourt. Golob, Sacha. 2014. Heidegger on Concepts, Freedom and Normativity. Cambridge: Cambridge University Press. Haugeland, John. 2000. “Truth and Finitude,” in Heidegger, Authenticity, and Modernity, edited by Mark Wrathall and Jeff Malpas (pp. 43–78). Cambridge: The MIT Press. ———. 2007. “Letting Be,” in Transcendental Heidegger, edited by Steven Crowell and Jeff Malpas (pp. 93–103). Stanford: Stanford University Press. Heidegger, Martin. 1962. Being and Time, translated by John Macquarrie and Edward Robinson. New York, NY: Harper & Row. ———. 1982. Basic Problems of Phenomenology, translated by Albert Hofstadter. Bloomington: Indiana University Press. ———. 1998. Pathmarks, edited by William McNeill. Cambridge: Cambridge University Press. Husserl, Edmund. 1989. Epilogue. In Ideas Pertaining to a Pure Phenomenology and to a Phenomenological Philosophy, Second Book, translated by R. Rojcewicz and A. Schuwer, Dordrecht: Kluwer.

Determinable Indeterminacy  147 ———. 1959. Erste Philosophie (1923/24). Zweiter Teil, edited by Rudolf Boehm, Husserliana VIII. Den Haag: Matinus Nijhoff. ———. 1973. Experience and Judgment, translated by James S. Churchill and Karl Ameriks. Evanston, IL: Northwestern University Press. ———. 1983. Ideas Pertaining to a Pure Phenomenology and to a Phenomenological Philosophy, First Book, translated by F. Kersten. The Hague: Martinus Nijhoff. ———. 1965. Philosophy as Rigorous Science. In Phenomenology and the Crisis of Philosophy, translated by Quentin Lauer. New York, NY: Harper & Row. ———. 1970. The Crisis of European Sciences and Transcendental Phenomenology, translated by David Carr. Evanston, IL: Northwestern University Press. ———. 2004. Wahrnehmung und Aufmerksamkeit. Texte aus dem Nachlass (1893– 1912), edited by Thomas Vongehr und Regula Giuliani, Husserliana XXXVIII. Dordrecht: Springer. Kant, Immanuel. 1987. Critique of Judgment, translated by Werner Pluhar. Indianapolis, IN: Hackett. Levinas, Emmanuel. 1969. Totality and Infinity, translated by Alphonso Lingis. Pittsburgh, PA: Duquesne University Press. Loidolt, Sophie. 2015. “Transzendentalphilosophie und Idealismus in der Phänomenologie.” Methodo. International Studies in Phenomenology and Philosophy 103–135. Merleau-Ponty, Maurice. 2012. Phenomenology of Perception, translated by Donald A. Landes. New York, NY: Routledge. Natanson, Maurice. 1968. Being-in-Reality. In Literature, Philosophy, and the Social Sciences. The Hague: Martinus Nijhoff. Rouse, Joseph. 2015. Articulating the World. Chicago: University of Chicago Press. ———. 2002. How Scientific Practices Matter. Chicago: University of Chicago Press. Sartre, Jean-Paul. 1956. Being and Nothingness, translated by Hazel Barnes. New York, NY: Washington Square Press. Smith, Arthur David. 2003. Husserl and the Cartesian Meditations. London: Routledge. Tengelyi, László. 2014. Welt und Unendlichkeit. Freiburg: Alber. Wrathall, Mark. 2005. “Motives, Reasons, and Causes,” in The Cambridge Companion to Merleau-Ponty, edited by Taylor Carman and Mark Hansen (pp. 111– 128). Cambridge: Cambridge University Press. Zahavi, Dan. 2018. Husserl’s Legacy: Phenomenology, Metaphysics, and Transcendental Philosophy. Oxford: Oxford University Press.

6 Climate Science, Indeterminacy, and Food Security in Sub-Saharan Africa Trish Glazebrook and Michael Goldsby

This chapter assesses the function of scientific indeterminacy in hampering climate justice for sub-Saharan Africans. Most vulnerable to climate impacts affecting food security are the world’s poor who rely on subsistence agriculture to feed their family. In sub-Saharan Africa, as across most of the global South, the majority of these farmers are women. In Ghana, for example, women grow 87 percent of what goes into the national food basket.1 Accordingly, at stake in the question of climate justice for the global South are principles of distributive justice concerning who suffers the harms and who benefits from climate change, food justice with respect to increasing global inequities in food access, gender justice because food security is largely women’s responsibility, and intergenerational justice insofar as nutritional deficiency increases infant mortality and has lifelong consequences for childhood development. Indeterminacy in climate science is a complex question that has been used to smokescreen these justice issues and to deter policy response to the urgent, global threat to ecosystems and species, including human being, posed by anthropogenic greenhouse gas emissions. Yet humans are engineering the next great extinction event—only five are previously evident in the fossil record—through the generation of greenhouse gases, and many are already suffering in consequence. The 5th Assessment Report (AR5) of the Intergovernmental Panel on Climate Change (IPCC) models its projections of future climate change using four Representative Concentration Pathways (RCPs) that differ in their assumptions about how much effort is made to cut the greenhouse gas emissions driving climate change. The worst-case scenario, RCP8.5, models climate change under a businessas-usual, no mitigation approach. More than half the outcomes that result when the model is run using different initial conditions or parameters show a Global Mean Surface Temperature (GMST) increase of more than 4˚C by the end of the 20th century.2 This means that failure to mitigate could very well make much of the planet uninhabitable for many species, including humans. Our goal is to show that indeterminacy in the probabilistic projections made by climate scientists does not justify stalling policy response to the changing climate.

Climate Science, Indeterminacy, and Food Security  149 We argue that such stalling is irresponsible, unethical and not epistemically rational. We begin with an explanation of why we focus on climate mitigation, and provide evidence that climate denial is hampering mitigation policy. Then we provide an overview of the current humanitarian hunger crisis in sub-Saharan Africa. Predictions for Asia are just as dire, but catastrophic food security breakdown is already widespread across the African continent. Next, we distinguish two uses of the term probability among climate scientists: one is an expression of uncertainty, the other an expression of indeterminacy. We argue that there is no rational basis for using either as an excuse for failing to devise and implement mitigation policy. In conclusion, we bring our arguments back to food security in Africa to provide ethical argument for mitigation policy in the global North. As a prefatory note, our methods may be tedious for some readers as an analytic assessment of climate modeling. Heidegger argues that the essence of technology “sets upon nature”3 to drive “the organized global conquest of the earth”4 in what ecofeminists diagnose as a “logic of domination.”5 In this logic, humans make “the unreasonable demand that [nature] supply energy that can be extracted and stored.”6 Glazebrook and Gessas argue that this logic is paradigmatic in the fossil fuel industry.7 Burning fossil fuels is the cause of anthropogenic climate change. Heidegger suggests that “essential reflection upon technology and decisive confrontation with it” can happen in art.8 Our method in this chapter is to explicate the methodology of modeling in climate science to demonstrate that neither uncertainty nor indeterminacy provides a rational basis for climate denial that might preserve the economic hegemony of fossil fuels. Thus, we demonstrate that not art but science itself in climate debates confronts Heidegger’s “essence of technology.”9

Mitigation and Climate Denial At the policy level, climate change poses two challenges. First, mitigation is the challenge of cutting greenhouse gas emissions caused by burning fossil fuels as an energy source. This means reducing emissions in the global North while supporting poverty-alleviating development in the global South so these countries do not follow the high-carbon development pathway of industrialization taken by the global North. Mitigation is, according to current science, the best chance for keeping the rise of the GMST to 2˚C rather than 4˚C. Secondly, adaptation is the challenge of adjusting to the changing conditions of a complex global environment undergoing temperature increase. Adaptation can lessen the harm caused by climate change, but only mitigation holds the possibility of slowing, stopping, or even reversing climate change. Mitigation addresses the causes of climate change, while adaptation provides only post hoc attempts to live with its consequences. Moreover, adaptations will never be able to balance an ever-increasing GMST.

150  Trish Glazebrook and Michael Goldsby Given current impacts, adaptation is crucial for alleviation of the human suffering caused by climate change, but our focus is on the mitigation of greenhouse gas emissions in the global North. We take this focus in response to Dale Jamieson’s suggestion that climate change calls for a new notion of responsibility not driven by the assumption of the locality of harms.10 That is, we take the global North to be morally responsible for preventing further harm to the people of the global South by taking action to slow, halt, and reverse climate change, so our argument is toward functional climate mitigation policy creation and implementation in the global North. The needs of people in the global South stand at the intersection of climate, food, and gender justice, and depend on acceptance by policy-makers that anthropogenic climate change is real. McCright and Dunlap (2000) identified three prevalent ways of dismissing the reality of climate change: criticism of scientific evidence, appeal to benefits from climate change, and the economic threat that action on climate change poses.11 The George W. Bush Administration continuously emphasized uncertainties in scientific forecasts of climate change as well as the economic impacts of mitigation policies.12 Senator James Inhofe, Chair of the Senate Committee on Environment and Public Works, told Congress in 2003, for example, that “all the phony science” was contributing the “greatest hoax ever perpetrated on the American people” that threatened to “destroy the foundation, the greatness of the most highly industrialized nation in the history of the world.”13 Antilla assessed climate science in news media on four aspects: validity, ambiguity of cause and effect, controversy, and uncertainty.14 These are not unrelated; for example, controversy is generated using the other kinds of arguments, e.g., that the science is not valid, that climate change could just as well be caused by something other than human activity, and that the science is uncertain. Most of the articles examined were found to accept the validity of science, though other studies found informational bias in 70 percent of news media examined, and as much space given to climate deniers as to scientists in half the specimens examined.15 Earlier, Boykoff attributed such “balanced” reporting to scientific uncertainty that gave news outlets opportunities to draw from a range of sources.16 Uncertainty remains as a long-standing, focal criticism of climate science and as a basis for non-action.17 Beyond the popular press, it figures strongly in policy. For example, Scott Pruitt, Administrator of the US Environmental Protection Agency, said that the degree of human contribution to climate change is uncertain, which makes it difficult to decide how policy-makers should respond.18

Food Security in Sub-Saharan Africa Daily living conditions for many in Africa are already significantly affected by climate change because of their close reliance on their local ecosystem to support their subsistence livelihood. Agriculture is a critically important

Climate Science, Indeterminacy, and Food Security  151 issue in climate change insofar as agricultural technology responds to global challenges in food security while exacerbating climate change.19 Distributive justice is breached when those who contribute fewer of the greenhouse gases driving climate change than others nonetheless suffer the consequences more. Of the estimated 570 million farms in the world, only 4 percent are in high-income countries, i.e., the global North.20 Approximately 9 percent are in sub-Saharan Africa, where more than 80 percent of farms are less than two hectares (five acres) in size.21 Many of the farmers who work this land are women subsistence farmers whose primary use of crops is to feed their family rather than to generate income by bringing their crop to market. These populations rely on their agriculture to meet their family’s daily living needs; they depend on their crops for survival. In sub-Saharan Africa, a massive humanitarian crisis in food security is already underway. In 2008, twenty-one of the thirty-six countries facing food insecurity were in Africa.22 That is to say, well over half the world’s countries that cannot adequately feed their people are on only one of the six permanently inhabited continents. This global disproportion has been steadily increasing: from 1990 to 2016, the percentage of the world’s undernourished living in sub-Saharan Africa went from 17 percent to 27 percent.23 Between 1990 and 2010, the number of African countries facing food crises doubled from twelve to twenty-four; nineteen of these countries experienced food crisis in eight or more years of the ten years between 2005 and 2015.24 Today, every fourth sub-Saharan African is undernourished,25 i.e., 220 million people on the continent, many of them children, do not have enough to eat.26 These hunger issues are affected by economic factors. In 2008, the United Nations’ Food and Agriculture Organization reported four African countries in acute food crisis because of supply bottlenecks, distribution complications, and limited imports. They also identify seventeen other African countries in which extreme poverty and population displacement contributed to local food shortages and more general food access insufficiency.27 The ratio of GDP growth against the number of farmers is globally the lowest in the South—meaning that gaps have grown and continue to grow between how much food is needed and how much crops yield. In 2002, more than 200 million sub-Saharan Africans were living on less than $1 USD per day.28 In 2012, 47 percent, i.e., more than 500 million people, lived on less than $2 USD per day.29 Many sub-Saharan farmers cannot afford energy-hungry technologies powered by fossil fuels, and typically farm with a handhoe. They also cannot afford agricultural extension services, such as fertilizers, or expensive inputs, such as genetically modified seeds. In recent years, the gap between food security requirements and agricultural yield has grown more in the global South than in the North. This poverty means that subsistence farmers in sub-Saharan Africa have a small carbon footprint because they cannot afford fossil-fuel burning technologies. At the same time, they are ill

152  Trish Glazebrook and Michael Goldsby prepared and under-resourced to respond and adapt to environmental factors that cause natural disasters leading to crop failure and large post-­ harvest losses.30 The context in which climate change challenges African food security is extremely complex, and is impacted by economic as well as environmental factors. Since growing food is largely women’s responsibility in the subsistence economies of African countries, disproportionate climate impacts on agriculture in the global South breach distributive principles in climate justice, gender justice, and food justice. The health consequences of nutritional deficiency for children further breach principles of intergenerational justice. Mitigation is called for in order to remedy these global injustices. The challenges of mitigation unfold at the intersection of science, policy, and capital. Scientists seek to understand climate change; policy-makers seek to respond by reducing emissions and building adaptation capacity; corporations, especially oil companies, seek protection for investments in fossil fuel powered production and infrastructure. Since the inception of the United Nations Framework Convention on Climate Change (UNFCCC), policy-makers have been caught in a stalemate between climate science and corporate lobbying. Public demand for action on climate change has moreover been stymied by an active, corporate-sponsored climate denial industry aimed at fostering popular skepticism concerning the reliability of climate science. There is good evidence that climate denial is intentionally manufactured by the same propaganda experts who worked for the tobacco industry to create skepticism concerning the health consequences of tobacco use.31 The next three sections refute climate denial arguments relying on claims that climate science is probabilistic.

Probability With respect to probabilistic approaches in climate science, we distinguish uncertainty from indeterminacy. Uncertainty, as we use the term in the context of climate science, is best understood as the difference between the degree of confidence scientists have in the models and theories of climate science, given the evidence, and absolute confidence in those models. Indeterminacy, on the other hand, is the result of mismatches, whether known or unknown, between climate models and the target system (i.e., the actual climate system). As a result of indeterminacy, climate scientists design their models to provide probabilistic predictions for a range of possible future states, rather than precise point predictions of the future climate that more deterministic models would provide. The IPCC’s AR5 adopted a Guidance Note for Lead Authors on Consistent Treatment of Uncertainties.32 The IPCC does not distinguish, as we do, between uncertainty and indeterminacy. Nonetheless, it has two metrics for communicating the degree of certainty based on the authoring teams’ evaluations of scientific understanding. The first metric, confidence in the validity

Climate Science, Indeterminacy, and Food Security  153 of a finding, measures climate science research based on the type, amount, quality and consistency of the evidence, as well as agreement within the climate science community. Confidence in the validity of a finding is a measure of both how well the finding is supported by the evidence as well as a measure of scientific consensus related to the finding, where the available evidence is described as limited, medium, or robust, and the degree of agreement is rated as low, medium, or high. This produces a confidence range going from very low confidence (low agreement across limited data) to very high confidence (high agreement across robust data), which can be expressed probabilistically. The second metric is a quantified measure of certainty concerning the likelihood of an outcome having occurred or occurring in the future. Assessment is probabilistic and ranges from a less than 1 percent probability labeled exceptionally unlikely to an above 99 percent likelihood labeled virtually certain. Part of what makes space for climate denial is that probability is a term used ambiguously by climate scientists, even when regimented by the IPCC. Goldsby and Koolage discuss two ways the term is used in model-based science, including climate science.33 The first concerns the probability that a certain outcome will obtain. For example, when climate modelers say in AR5 that there is a greater than 66 percent chance that under a particular scenario the GMST will increase between 2.6˚C and 4.8˚C by 2100, they are reporting that more than 66 percent of the model’s projections indicate an increase within that temperature range. There are many factors at work in climate change, just one of which is the fact that a perfectly accurate picture of initial conditions is not available. Measurement errors and unknown factors affecting the system are just a couple of sources of that inaccuracy. To deal with this issue, modelers use several runs of a model, wherein they may vary the initial conditions or the parameters of the model, in order to provide a range of possible and likely outcomes. When this is done, more than two-thirds of the outcomes show a temperature increase between 2.6˚C and 4.8˚C by 2100.34 This first sense of probability simply refers to outputs of a model based upon the frequency of a projected outcome. This is the modeloutput probability, and it is best seen as an expression of the indeterminacy that exists within climate science. A second use of probability is an expression of how confident one is in the models and theories used, given the evidence, and consequently how confident one is in the results, i.e., the probabilistic predictions and projections derived from those models and theories. This probability is best called a credence, to use the common Bayesian term. According to Bayesianism, epistemic certainty or credence can be expressed as some probability p such that 0≤p≤1 (0 or 100 percent or any probability in between). For those claims about which there is absolute certainty (e.g., tautologies), the credence should be 1 (100 percent), and 0 (0 percent) is reserved for those claims about which we can be absolutely certain that they are false (e.g., contradictions). Contingent claims can be assigned any probability between

154  Trish Glazebrook and Michael Goldsby 0 and 1, but should be updated higher toward 1 when confirming evidence is discovered (and the degree of uncertainty is reduced), and lower when one discovers disconfirming evidence. The more evidence one obtains in favor of a theory, the higher the credence should be, and the more evidence against the theory, the lower the credence should be. High confidence in a theory suggests that the credence is very close to 1, and that there is considerable evidence in favor of the theory. Furthermore, when one bases one’s credence on the evidence, then the relationship between confidence and uncertainty will be an inverse relationship, that is, higher credences will indicate lower levels of uncertainty. That is the case with anthropogenic climate change. For example, AR5 states that the IPCC has very high confidence that anthropogenic CO2 emissions are the largest contributor to warming trends. Similarly, the IPCC reports that it is virtually certain that warmer days will be more frequent and cooler days less frequent as the 21st century comes to a close. With regard to models, the IPCC reports very high confidence in the model’s ability to predict and retrodict global features of the climate such as GMST.35 As Goldsby and Koolage point out, it is the second probability that “is doing the epistemic heavy lifting” in climate science.36 In other words, these expressions of confidence are probabilistic expressions that encode that there is plenty of evidence to support that climate change is real, anthropogenic, and likely to be bad if not addressed. The next two sections make our arguments against climate denial based on these two meanings of probability.

Uncertainty Even very high confidence or virtually certain are far enough away from certainty that rival models and theories that might suggest contrary results cannot be absolutely ruled out. In fact, a climate denial theory that is inconsistent with the best contemporary climate models may in principle be “held true come what may,”37 even against seemingly contradictory evidence, provided that one is willing to make enough changes in the auxiliary assumptions that support the theory. Quine made this point in response to Duhem’s argument that physical theories could not generate testable predictions without the addition of auxiliary assumptions; a failed prediction in a physical experiment accordingly indicates that either the hypothesis or an auxiliary assumption is wrong, but not which.38 Quine extended Duhem’s results to the whole of science by maintaining that any theory could be saved from inconsistency with recalcitrant data if severe enough changes were made to the auxiliary assumptions.39 Quine’s point is a logical one— namely that mere consistency with empirical evidence is not sufficient to determine which theory is the best theory. A climate denier might misappropriate Quine’s thesis and deny that it is reasonable to believe that climate change is happening, that it is the result of human activity, or that it represents a serious threat to humanity. Rival models cannot simply be ruled

Climate Science, Indeterminacy, and Food Security  155 out as there are so many auxiliary assumptions that might be changed. In what follows, we will argue that while consistency with empirical evidence is necessary for rationality it is not sufficient, and that while climate deniers might meet that minimum benchmark, they nonetheless fall short. For example, perhaps the most accessible piece of evidence that the climate is changing is the fact that the GMST has climbed 0.8˚C (1.44˚F) over the past one hundred and thirty years that it has been directly measured across the globe. Since in a sense global warming is synonymous with rises in GMST, it would seem hard for deniers to hold onto their theories that climate change is not happening, in light of this evidence. Nonetheless, climate deniers might insist that the record of temperature rise is either biased or unreliable (i.e., they might deny the assumption that the data are veritable). McKitrick and Michaels tried to do just that by claiming that the “apparent” temperature rise is merely the result of increasing urbanization and the phenomenon of urban heat islands.40 Even when it is shown that rural temperatures have increased at approximately the same rate as urban temperatures, climate deniers can still hold on to the rival hypothesis by questioning whether 19th-century temperature readings are accurate, or even sowing doubt about the honesty of the scientists compiling the record. The rival denialist theory can never be completely eliminated from a logical perspective—no matter how much climate scientists play whack-a-mole— because either new assumptions can be added (e.g., scientists are dishonest) or reasonable assumptions might be questioned (e.g., whether 19th-century measurements are accurate). Furthermore, since science is not practiced in isolation, there will always be assumptions that can be questioned, added, or modified to save the denier’s theory that climate change is not happening. There will always be some uncertainty insofar as any rival theory can adjust assumptions to make the evidence consistent with the theory. As Laudan points out, however, even if such rival theories cannot be ruled out because they can be made consistent with any empirical evidence, this does not mean that one can rationally hold such a theory against mounting, contrary evidence.41 It is, moreover, an open question whether such denial theories can be made consistent with empirical data in a non-trivial way.42 Furthermore, the adjustments to the set of auxiliary assumptions necessary to make climate denial theories consistent with empirical evidence may produce models, the predictions of which cannot be trusted in making policy decisions.43 Finally, the demand for certainty is “an interesting theoretical exercise in epistemology, but [it] ought not to be applied selectively against climate science.”44 If absolute certainty were the standard for accepting scientific claims, nearly every scientific claim—from evidence-based medicine to meteorology to epidemiology and even astronomical models of the solar system—would have to be rejected. Selectively applying the absolute-­ certainty standard to climate science “suggests a worrying kind of epistemic and moral corruption.”45 One of the strengths of science is that any theory

156  Trish Glazebrook and Michael Goldsby is revisable in the face of more evidence. Acknowledging some uncertainty in even the most well-supported theories is the only way to take advantage of that strength. After all, those things about which certainty is absolute cannot be revised. Hence the IPCC provides metrics assessing its findings— not because the core claims of climate science are not well-supported by the evidence, but rather as an indication and acknowledgement that there is still more to learn. Still, it should be stated that if scientific confidence in the reality of climate change ranges from “very high” to “virtually certain,” then that suggests that it is irrational to place much confidence (if any at all) in denialist theories. Appeals to uncertainty are thus disingenuous at best, and at worst dangerously misguided. As we have shown, these appeals to uncertainty are not rationally grounded, and as such ought not be used to provide political cover for ignoring the consequences of climate change already being felt throughout the world, but especially in Africa’s humanitarian crisis in food security. David Malone notes: As every climate scientist knows, there will always be facts that won’t fit even the best model of global climate. That’s the nature of models and the weather—and it illustrates just how badly we can be led astray by the fiction that science is about certainty. If we are honest and say the scientists’ conclusions aren’t certain, we may find this being used as justification for doing nothing.46

Indeterminacy Appeals to indeterminacy are another matter. Whereas uncertainty is a measure of how far our credence in theories departs from absolute certainty, indeterminacy relates more to model-output probabilities. To better illustrate this point, consider initial condition sensitivity. Given that observations are rarely, if ever, perfect, error tends to creep in when determining what the current conditions of a complex system are. Those errors are typically tiny; however, tiny errors can be magnified and amplified over a series of time steps such that radical departures from the evolution of the true system will result. This is the so-called butterfly effect. Nonetheless, scientists have learned to manage this problem. One way that they do so is to assume that there is some error in their observation of current conditions, and rather than use their point measurements as the input for their models, they instead input a probability distribution of what the current conditions might actually be. As a result, the output of such model-runs is also a probability—a model-output probability. This strategy tends to ameliorate the butterfly effect, producing reliable, but probabilistic predictions and projections. Deniers’ appeals to indeterminacy have both naïve and sophisticated versions. The naïve version conflates or equates indeterminacy with uncertainty. That, however, is a mistake. It is true that in some cases, uncertainty

Climate Science, Indeterminacy, and Food Security  157 can result in indeterminacy. For example, indeterminacy arising from initial condition sensitivity does find its origin in uncertainty about initial conditions. Nonetheless, indeterminacy is distinct from uncertainty. To illustrate that point, consider weather models. A weather model in which there is high confidence, i.e., the certainty measure is greater than 90 percent credence that the model provides good probabilistic predictions, may provide a mere 15 percent chance of rain as a model output. Nevertheless, one would be reasonable in having a high confidence that it is unlikely to rain. Likewise, casinos and insurance companies can have high confidence that they will make money in the long run, even though there is indeterminacy as to whether they will turn a profit next week. Similar claims could be made about evidence-based medicine, epidemiology, meteorology, and all sciences that deal with probabilistic models. Indeterminacy is accordingly not necessarily a sign of uncertainty; inevitability of indeterminacy provides no more rational basis for denying empirically well-supported climate models than mere uncertainty does. Most contemporary science is model based.47 Indeterminacy is part and parcel of model-based science because models predict future states of a system using data known about its current state in combination with an understanding of the factors that change the system, expressed for example in a transformation equation. Setting aside issues of whether the world is fundamentally deterministic or stochastic, indeterminacy usually results from mismatch between a model and its target system that is the result of idealizations used in model construction. Common idealizations in climate models include omissions (either intentional or unintentional), parameterization (the use of a single parameter to represent a process that is far more complex), or the use of coupling mechanisms in complex models that allow modelers to use more manageable sub-models to model a larger more complex system. These strategies tend commonly to omit some of the target system’s complexity. Climate deniers may, nonetheless, marshal a more sophisticated critique by citing these sources of indeterminacy. After all, mismatches between model and target ­system—in this case the global climate system—raise the question as to just how reliable the model’s predictions might be. To respond to that challenge, we turn to a field of statistics known as model selection theory. Consider cases where climate scientists intentionally leave out some causal factors known to affect the target system. Examples include (but are not limited to) airplane contrails,48 and wildlife methane emissions.49 While both are known to have some impact on climate, no global climate model takes them into account. Would not a global climate model that accounts for these factors provide better predictions of the state of the future climate? Surprisingly, accounting for these factors may actually make predictions worse. The counterintuitive result from model selection theory is that false or incomplete models sometimes provide better predictions than true and complete ones.50

158  Trish Glazebrook and Michael Goldsby In order to illustrate how this can be the case, consider an illustrative example co-opted from Sober of curve-fitting, i.e. drawing a line to capture the trend across several data points.51 Suppose we want to model (for predictive purposes) the relationship between temperature and pressure in a pressure cooker. The first thing we would want to do is gather some temperature and pressure measurements. Of course, we are not measuring those features so much as taking readings from generally reliable but not perfect measuring devices. After we gain a couple of dozen measurements, we plot them on a graph depicted in Figure 6.1a and 6.1b. For the purpose of this illustration, let’s suppose that we are considering two different models for the relationship between temperature and pressure. The first is labeled LIN because it expresses a linear relation, i.e., a relation between two variables that can be graphed as a straight line. The second is labeled 1PPDP because it plots one parameter per data point. The first can be described by a simple equation:

(LIN)

y = a + bx + e

According to LIN, the relationship between the temperature (x) and pressure (y) is linear. LIN is an unfitted model, i.e. the values of its three adjustable parameters: a, b, and an error term e have not yet been determined. A fitted model, in contrast, is one where all the adjustable parameters are replaced with fixed values, through a part of the modeling process known as parameter estimation or fitting. As such, we might think of LIN as a family of fitted models. To fit or train the model to the data we collected, we must estimate what the parameters are. One way of doing so is to determine L(LIN), i.e. the fitted model of LIN with the highest likelihood. The line in Figure 6.1a represents L(LIN). One thing to notice about L(LIN) is that while it goes through a few data points, there are several that are not on the line. Additionally, there are at least a couple of outliers that fall quite far from the line. Can we improve how well our model fits the data? The answer is yes, but only if we add more parameters and consequently more complexity. LIN is perhaps the simplest model we can use to model the relationship. Now let’s consider another model represented in Figure 6.1b whereby we maximize the fit-to-data, by using one parameter for each data point (1PPDP):

(1PPDP)

y = a + bx + cx 2 + dx3 + fx 4 + ... + wx 21 + zx 22 + e

There are a couple of things to notice about 1PPDP. First of all, 1PPDP is ridiculously complex. It is a 22nd order polynomial with 24 adjustable parameters—one for each data point. Secondly, given the complexity, it is a trivial—albeit onerous—mathematical task to fit the model such that the resulting curve passes through each and every data point.

Climate Science, Indeterminacy, and Food Security  159 (a)

Pressure

Pressure

(b)

Temperature

Temperature

  Figure 6.1  These graphs plot the same observed temperature and pressure readings in our pressure cooker example. 6.1a shows the best fitting model L(LIN), while 6.1b represents the best fit-to-data.

Scientists rarely construct such models because models of commensurate complexity are rarely necessary for the work they do. More importantly there is a reason that they ought not use such maximally complex models. More complex models fit the data already obtained much better than simpler models, but they often do worse at predicting new data. Those models that have one adjustable parameter per data point do the worst of all, when fit to the training data set. The problem is that maximally complex models tend to overfit the data, effectively treating every data point as perfectly accurate, and as such it treats outliers as causally relevant and not as possible mistakes in measurement (see Figure 6.1a and 6.1b). In other words, a model that overfits the data tends to mistake noise (e.g. errors in measurement) in data collection for the signal of true causal features in the target system.52 As a result, if we were to collect new measurements, those new measurements would be far more likely to fall close to L(LIN) than the fitted 1PPDP model. This does not mean that the simplest model is always the best predictor. Sometimes, the increased fit of a more complex model offsets problems with increasing the complexity of the model. After all there are plenty of other models between 1st-order and 22nd-order polynomials. The goal of model selection theory is to manage best the tradeoff between simplicity, i.e. minimizing adjustable parameters, and fit-to-data. Hirotugu Akaike (1973) showed that a modeler can maximize predictive accuracy by optimizing this tradeoff. In layperson’s terms,53 models are compared to see how well they fit the data at hand, and then each model is assessed a penalty for each added parameter. To maximize predictive accuracy, one must ask whether increased fit through adding another parameter offsets or exceeds the penalty. The answer depends upon the size and character of the data set used to train the model. With smaller data sets, simpler models tend to be better predictors; with larger data sets, more complex models tend to do better if increased data reveal increased complexity.54

160  Trish Glazebrook and Michael Goldsby Returning to climate change, why do climate modelers leave out some features, such as airplane contrails and wildlife methane emissions? The answer is that there are some factors that have such little impact that explicitly accounting for those features in a global climate model increases complexity without a commensurate increase in our ability to represent the system. Adding in factors for the methane emissions of wild game and accounting for the albedo effect of contrails has such a little impact on the predictive capabilities of the models that it does not offset the penalty for increased complexity. Climate modelers use model selection tools like those developed by Akaike to maximize the predictive accuracy of their models, and intentional omissions must sometimes be made. So climate deniers, in order to make their case, would have to show that the omitted features increase the model’s fit without overfitting the data. Since no climate denier has provided a model shown to be better at optimizing predictive accuracy, the climate denier has no rational basis for denying the conclusions of IPCC Assessments Reports. Deniers thus have no rational justification for hampering mitigation policy by refusing to accept the science.

Conclusion Laudan attempts to debunk the Duhem-Quine thesis discussed earlier because he rejects the claim that acceptance of scientific theories hinges in part on non-scientific factors.55 This may be true for scientists. But when it comes to public discussion of climate science by non-experts, Lemons et al. argue that the climate debate is not so much about science as it is about values: it has been disguised as a discussion about epistemology, e.g., our account of climate modeling in science, while “it would be better to discuss our ethical differences explicitly and directly rather than to mask them in the language of science.”56 This “mask” is, however, typically engaged by climate deniers: instead of confronting the ethical questions raised by climate change, they put into question epistemology, i.e., the status of science as knowledge, in order to provide cover for their reluctance to act. Jamieson argues that climate change engages fundamental questions “about how we ought to live, what kinds of societies we want, and how we should relate to nature and other forms of life.”57 He calls for a new value system that entails a new notion of responsibility that does not assume that “harms and their causes are individual, that they can be readily identified, and that they are local in time and space.”58 Our intent has been precisely to create a space for ethical responsibility that avoids these assumptions and obliges action by the global North to prevent further harms to people living in the global South from climate impacts largely caused by the global North. We have done this by showing that indeterminacy in climate science provides no epistemological basis for climate denial that might justify refusal to mitigate anthropogenic greenhouse gas emissions.

Climate Science, Indeterminacy, and Food Security  161 There is a further philosophical point to be made here. To paint with broad strokes, hermeneutics originated as interpretive theory concerning biblical texts, expanded to the interpretability of any text, and grew with Heidegger’s Sein und Zeit into the insight that all human experience is situated, interpretive praxis. Therefore, rather than truth being universal, eternal, and unchanging, there can be different “truths,” i.e., differing accounts of reality, situated in different cultural traditions or historical epochs. For example, Aristotle held that fundamentally earthly things are drawn to the center, Newton that the earth exerts a gravitational force. Both are good reasons not to drop your iPhone; but Newton’s account is better if you want to design a rocket to take you to the moon. From this view of human experience as inherently hermeneutic has emerged a contemporary understanding in popular culture that “postmodernism” is the belief that anything goes with respect to truth. Our analysis of climate modeling provides a concrete example of how a system can be understood to encompass many diverse outcomes, i.e., a multiple rather than singular actuality, without it being the case that any outcome is possible. The issue is not what the truth is, but what it will be as climate change unfolds. Moreover, it is possible to make best-case assessments, where “best-case” means not the most favorable outcome but what is most likely to hold. So, while probabilistic modeling in climate science takes into account an element of indeterminacy, not just anything goes, and among what does go, there are better or worse options upon which to base policy. From the core of what in contemporary science is most decisive for the future of the planet has thus emerged a reconciliation between analytic, Anglo-American philosophy and continental philosophy’s repudiation of any realism that assumes the existence of a single reality that might ultimately be exhaustively described by a single, correct account. Rather, our analysis of climate modeling demonstrates that reality is best thought as a complex arrangement of better or worse practical possibilities. This is precisely the Heideggerian point that “higher than actuality stands possibility.”59 Heidegger also argued that the ability to manipulate nature toward realizing envisioned possibilities is destructive when nature is ontologically reduced to instrumental value.60 Science is not just theory but has practical consequences in that scientific knowledge drives technological applications. Research funding agencies not only understand that science has practical consequences, but demand that scientists also understand and articulate the “broader impacts” or “outputs and outcomes” of their work. As Hacking argued, science not only represents reality, but also intervenes.61 Not all of the consequences of human activity are intentional. Climate change is an unintended consequence. Climate deniers are reneging on their moral responsibility to be accountable for the consequences of their activities, and are moreover asking their governments to renege on their responsibility to participate in global policy concerning climate mitigation and adaptation.

162  Trish Glazebrook and Michael Goldsby One consequence of global climate change is that in Africa, a humanitarian crisis in food security is well underway, yet only getting started. AR5 notes that marginalized groups are especially vulnerable to climate change that exacerbates other livelihood stressors, e.g., decreasing crop yield, for people living in poverty.62 They have sparse resources to fall back on when yield falls sort, and since only 3.7 percent of arable land in Africa is irrigated (NEPAD 2002), are particularly vulnerable to changing rainfall patterns and drought. AR5 projects with high confidence that if the GMST increases by 2˚ C, African crop productivity will be nearing medium risk by 2030 and well into it by 2080, even with adaptation; models show very high risk to African agriculture for a 4˚C rise.63 Distributive principles in environmental justice with respect to climate, gender, food, and intergenerational justice are accordingly breached, in that those generating comparatively few of the greenhouse gases driving global climate change are experiencing life-threatening food insecurity, while those in countries where most greenhouse gases are generated for the most part have the resources to buy food and in other ways respond to climate impacts. Failure to take strong and immediate action on climate mitigation in the global North shows indifference to human suffering and life in the global South. We find that such inaction is irresponsible and unethical, and attempts to deter action through appeal to the uncertainty or indeterminacy of climate science are inexcusable.

Notes 1 Social Watch Coalition (2010). 2 IPCC (2013, p. 1055). 3 Heidegger (1997a, p. 18). 4 Heidegger (1997b, p. 358). 5 Warren (1996, pp. 21–24). 6 Heidegger (1997a, p. 18). 7 Trish Glazebrook and Jeff Gessas (in press). 8 Heidegger (1997a, p. 39). 9 Cf. Trish Glazebrook (in press). 10 Jamieson (2014, p. 292). 11 McCright and Dunlap (2000). 12 Jamieson (1992, p. 141). 13 Inhofe (2003, S10022). 14 Antilla (2005, p. 344). 15 Boykoff (2008). 16 Boykoff (2007, p. 108). 17 Antilla (2005, pp. 345–347 and 350). 18 Leber and Schulman (2017). 19 Glazebrook (2017, pp. 304–305). 20 Lowder et al. (2016, pp. 20–21). 21 iIbid., p. 25. 22 Kabasa and Sage (2009, p. 21). 23 FAO (2015, p. 10).

Climate Science, Indeterminacy, and Food Security  163 4 Ibid., p. 27. 2 25 Ibid., p. 12. 26 Ibid., p. 8. 27 FAO (2008). 28 World Bank (2007). 29 WHES (2016). 30 FAO (2008). 31 Monbiot (2006) and Oreskes and Conway (2010). 32 IPCC (2013, p. 36). 33 Goldsby and Koolage (2015). 34 IPCC (2013, p. 23). 35 Ibid., p. 75. 36 Goldsby and Koolage (2015, p. 231). 37 Quine (1951, p. 51). 38 Duhem (1954). 39 Quine (1951, p. 40). 40 McKitrick and Michaels (2007, S09). 41 Laudan (1990). 42 Cf. Grünbaum (1974). 43 Cf. Goldsby and Koolage (2015). 44 Gardiner (2011, p. 173). 45 ibid. p. 463. 46 Malone (2007, p. 47). 47 Cf. Giere 1988 and Suppe (1977). 48 Lee et al. (2009). 49 Röös et al. (2014). 50 Forster and Sober (1994). 51 Sober (2008). 52 Cf. Forster and Sober (1994), Forster and Elliott Sober (2010), Burnham and Anderson (2002), and Goldsby (2013). 53 Hirotugu Akaike (1973). For those who prefer a more technical exposition, Akaike’s theorem provides a framework from which to compare the estimated predictive accuracy of two or more models. This framework uses the Akaike information criterion (AIC). The AIC score of a model M, AIC(M) =def log{Pr[data|L(M)]}—k. The term, L(M), refers to the fitted model of M that has the highest likelihood of all possible fitted models of M when compared to the training data, and Pr[data|L(M)] is that likelihood. The complexity of the model in terms of adjustable parameters is represented by k. The log-likelihood of L(M) is a measure of how well the model fits the data and k represents a correction for complexity. AIC(M) provides an unbiased estimate of M’s predictive accuracy—where predictive accuracy is defined as how well the model predicts new data after it has been fitted to old data. See Forster and Sober (1994), Forster (2000) and (2001), and Sober (2008). The absolute value of the score is not important; what is important is that AIC can be used to compare the estimated predictive accuracy of unfitted models, such that the model with the higher AIC score would better predict new data. See Goldsby (2013). 54 Goldsby (2013). 55 Laudan (1990). 56 Lemons et al. (1995, p. 132). 57 Jamieson (1992, p. 147). 58 Jamieson (2014, p. 292).

164  Trish Glazebrook and Michael Goldsby 9 Heidegger (1986 [1927], p. 38). 5 60 Heidegger (1997a; cf. Glazebrook [2000, p. 113]). 61 Hacking (1983). 62 IPCC (2014, pp. 6–8). 63 Ibid., p. 21.

Works Cited Akaike, Hirotugu. 1973. “Information Theory as an Extension of the Maximum Likelihood Principle,” in Second International Symposium on Information Theory, edited by B. Petrov and F. Caski (pp. 267–281). Budapest: Akademiai Kiado. Antilla, Liisa. 2005. “Climate of Scepticism: US Newspaper Coverage of the Science of Climate Change.” Global Environmental Change 15: 338–352. Boykoff, Max. 2007. “Mass Media and Environmental Politics,” in The Politics of the Environment: A Survey, edited by C. Okereke (pp. 101–115). London: Routledge. ———. 2008. “Lost in Translation? United States Television News Coverage of Anthropogenic Climate Change, 1995–2004.” Climatic Change 86(1–2): 1–11. Burnham, Kenneth P. and David R. Anderson. 2002. Model Selection and Multi-Model Inference: A Practical Theoretic-Information Approach. New York, NY: Springer. Duhem, Pierre. 1954. The Aim and Structure of Physical Theory, translated by Philip Wiener. Princeton: Princeton University Press. FAO. 2008. “Crop Prospects and Food Situation: Countries in Crisis Requiring External Assistance.” Food and Agriculture Organization of the United Nations. No. 2: April 2008. Online at www.fao.org/docrep/010/ai465e/ ai465e02.htm FAO. 2015. “International Fund for Agricultural Development, and World Food Programme. The State of Food Insecurity in the World 2015: Meeting the 2015 International Hunger Targets: Taking Stock of Uneven Progress.” Food and Agriculture Organization of the United Nations. Rome: FAO. Online at www.fao. org/3 /a4ef2d16–70a7–460a-a9ac-2a65a533269a/i4646e.pdf Forster, Malcolm. 2000. “Key Concepts in Models Selection: Performance and Generalizability.” Journal of Mathematical Psychology 44: 205–231. ———. 2001. “The New Science of Simplicity,” in Simplicity, Inference, and Modelling, edited by Arnold Zellner, H. A. Keuzenkamp, and Michael McAleer (pp. 83–119). Cambridge: Cambridge University Press. Forster, Malcolm and Elliott Sober. 1994. “How to Tell When Simpler, More Unified, or Less ad hoc Theories Will Provide More Accurate Predictions.” British Journal for the Philosophy of Science 45: 1–35. ———. 2010. “AIC Scores as Evidence—a Bayesian Interpretation,” in The Philosophy of Statistics. Dordrecht: Kluwer. Gardiner, Stephen. 2011. A Perfect Moral Storm: The Ethical Tragedy of Climate Change. Oxford: Oxford University Press. Giere, Ronald. 1988. Explaining Science: A Cognitive Approach. Chicago: University of Chicago Press. Glazebrook, Trish. 2000. “From physis to Nature, technê to Technology: Heidegger on Aristotle, Galileo and Newton.” The Southern Journal of Philosophy 38(1): 95–118.

Climate Science, Indeterminacy, and Food Security  165 ———. 2010. “Gender and Climate Change: An Environmental Justice Perspective,” in Heidegger and Climate Change, edited by Ruth Irwin (pp. 162–182). London: Continuum. ———. 2017. “Technology, the Environment, and Sustainability,” in Philosophy: Technology. MacMillan Interdisciplinary Handbooks: Philosophy series (pp. 293–320). Farmington Hills, MI: Macmillan Reference USA/Gale, a Cengage Company. ———. in press. “Letting Beings Be: An Ecofeminist Reading of Gestell, Gelassenheit and Sustainability,” in Heidegger and Technology, edited by Aaron Wendland. London: Routledge. Glazebrook, Trish and Jeff Gessas. in press. “Traditional Ecological Knowledge, Indigenous Rights, and Designing Water Policy,” in The Phenomenology of Water: Toward a New Policy Paradigm, edited by Ingrid Stefanovic. Toronto: University of Toronto Press. Goldsby, Michael. 2013. “The ‘Structure’ of the ‘Strategy’: Looking at the ­Matthewson-Weisberg Tradeoff and Its Justificatory Role for the Multiple-models Approach.” Philosophy of Science 80(5): 862–873. Goldsby, Michael and W. John Koolage. 2015. “Climate Modeling: Comments on Coincidence, Conspiracy, and Climate Denial.” Environmental Philosophy 12(2): 221–252. Grünbaum, Adolf. 1974. Philosophical Problems of Space and Time. Dordrecht: Reidel. Hacking, Ian. 1983. Representing and Intervening: Introductory Topics in the Philosophy of Natural Science. Cambridge: Cambridge University Press. Heidegger, Martin. 1986 [1927]. Sein und Zeit. 16. Auflage. Tübingen: Max Niemeyer Verlag. ———. 1997a. “Die Frage nach der Technik,” in Vorträge und Aufsätze. 8. Auflage (pp. 9–40). Stuttgart: Verlag Günther Neske. ———. 1997b. Nietzsche II, Gesamtausgabe, Band 6.2. Frankfurt am Main: Vittorio Klostermann. Inhofe, James. 2003. “Science of Climate Change.” Congressional Record 149(113): S10012–S10023. IPCC. 2007. Climate Change 2007: Synthesis Report. Summary for Policymakers. Geneva, Switzerland: IPCC. (AR4). ———. 2013. “Climate Change 2013: The Physical Science Basis,” in Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, edited by Stocker, T. F., D. Qin, G.-K. Plattner, M. Tignor, S. K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P. M. Midgley. Cambridge: Cambridge University Press. ———. 2014. “Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects,” in Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, edited by C. B. Field, V. R. Barros, D. J. Dokken, K. J. Mach, M. D. Mastrandrea, T. E. Bilir, M. Chatterjee, K. L. Ebi, Y. O. Estrada, R. C. Genova, B. Girma, E. S. Kissel, A. N. Levy, S. MacCracken, P. R. Mastrandrea, and L. L. White. Cambridge: Cambridge University Press. Jamieson, Dale. 1992. “Ethics, Public Policy, and Global Warming.” Science, Technology, & Human Values 17(2): 139–153. ———. 2014. Reason in a Dark Time. Oxford: Oxford University Press.

166  Trish Glazebrook and Michael Goldsby Kabasa, John David, and Irene Sage. 2009. “Climate Change and Food Security in Africa,” in Climate Change in Africa: Adaptation, Mitigation and Governance Challenges. (CIGI Special Report), edited by Hany Besada and Nelson K. Sewankambo (pp. 21–25). Waterloo, Canada: Centre for International Governance Innovation (CIGI). Laudan, Larry. 1990. “Demystifying Underdetermination,” in Minnesota Studies in the Philosophy of Science, Vol. XIV, edited by C. Savage (pp. 267–297). Minneapolis: University of Minnesota Press,. Leber, Rebecca and Jeremy Schulman. 2017. “Yes, the Mainstream Media Does Publish Fake News: A Time of Global Warming Denial in the Media. Mother Jones. Online July 5 2017 at www.motherjones.com/environment/2017/07/ timeline-climate-denial-news/ Lee, David S., David W. Fahey, Piers M. Forster, Peter J. Newton, Ron C. Wit, Ling L. Lim, and Robert Sausen. 2009. “Aviation and Global Climate Change in the 21st Century.” Atmospheric Environment 43(22): 3520–3537. Lemons, John, Rudolph Heredia, Dale Jamieson, and Clive Splash. 1995. “Climate Change and Sustainable Development,” in Sustainable Development: Science, Ethics, and Public Policy, edited by John Lemons and Donald A. Brown. Dordrecht: Springer-Science+Business Media, B. V. Lowder, Sarah K., Jakob Skoet and Terri Raney. 2016. “The Number, Size, and Distribution of Farms, Smallholder Farms, and Family Farms Worldwide.” World Development 87: 16–29. Malone, David. 2007. “Perspectives: Are We Still Addicted to Certainty?” New Scientist 2615: 46–47. McCright, Aaron M. and Riley E. Dunlap, 2000. “Challenging Global Warming as a Social Problem: An Analysis of the Conservative Movement’s Counter-claims.” Social Problems 47(4): 499–522. McKitrick, Ross R. and Patrick J. Michaels. 2007. “Quantifying the Influence of Anthropogenic Surface Processes and Inhomogeneities on Gridded Global Climate Data.“ Journal of Geophysical Research 112(D24): S09. Monbiot, George. 2006. Heat: How to Stop the Planet Burning. London: Penguin Books. NEPAD. 2002. Comprehensive Africa Agriculture Development Programme. Chapter 2 (2.1 and 2.2). Online at www.fao.org/docrep/005/y6831e/y6831e-03. htm#P442_104688 Oreskes, Naomi and Erik M. Conway. 2010. Merchants of Doubt. New York, NY: Bloomsbury Press. Quine, Willard Van Orman. 1951. “Two Dogmas of Empiricism.” Philosophical Review 60: 20–43. Röös, Elin, Cecilia Sundberg, and Peter-Anders Hansson. 2014. “Carbon Footprint of Food Products,” in Assessment of Carbon Footprint in Different Industrial Sectors, Volume 1, edited by Subramanian Senthikannan Muthu (pp. 85–112). Singapore: Springer. Sober, Elliott. 2008. Evidence and Evolution. Cambridge: Cambridge University Press. Social Watch Coalition. 2010. “MDGs Remain Elusive.” Social Watch: Poverty Eradication and Gender Justice. Online at www.socialwatch.org/node/12082 Suppe, Frederick. 1977. The Structure of Scientific Theories, 2nd ed. Champaign, IL: University of Illinois Press.

Climate Science, Indeterminacy, and Food Security  167 Warren, Karen. 1996. “The Power and Promise of Ecofeminism.” in Ecological Feminist Philosophies, edited by Karen Warren (pp. 19–41). South Bend: Indiana University Press. WHES (World Hunger and Education Service). 2016. “Hunger Notes: Africa Hunger and Poverty Facts.” Washington, DC. Online at www.worldhunger.org/ africa-hunger-and-poverty-facts/ World Bank. 2007. “World Development Report 2008: Agriculture for Development.” Washington, DC: World Bank. Online at http://siteresources.worldbank. org/INTWDRS/Resources/477365-1327599046334/8394679-1327606607122/ WDR_00_book.pdf

7 Genetic Phenomenology and the Indeterminacy of Racism Janet Donohoe

We frequently excuse a person’s allegiance to unacceptable views by referencing that person’s upbringing, his or her generation, or particular cultural influences. We acknowledge, for instance, that Western whites don’t understand their own white privilege, simply taking it for granted because it has been so much a part of Western culture. Perhaps we even uncomfortably allow certain slurs coming from grandparents because things were different in their era. Such actions reflect a conception of the sedimentation of values and ideas that bears investigation. In the following, I would like to suggest that deeply held views are part of the indeterminate horizon of every person’s constitution of the world. Some of those views can be made more determinate through encounters with what we consider to be alien or other. I draw upon Husserlian phenomenology to lay out the connection between indeterminacy and horizons. I examine the genetic phenomenological method of Rückfrage as a way of investigating the role of homeworld in our everyday experience. Homeworld serves as the foundation for our experience of what is “normal” and is significant in fashioning our style of constitution in the world. This has ramifications for how we understand any encounter with that which is alien or not home. Husserl calls us to be attentive to this relationship between homeworld and alienworld and encourages a position of renewal and critique that allows us to recognize the indeterminacy of homeworld and the possibility of thinking anew our homeworld traditions. I explain the genetic phenomenological approach, then explore what it means to address homeworld through the lens of genetic phenomenology and how that might be motivated by the encounter with alienworld. Finally, I explore the sedimentation of perceptions of race and how the hard investigative work of genetic phenomenology puts us in the position of renewal and critique that has ethical significance.

Husserl and Indeterminacy In his posthumously published work Crisis of the European Sciences and Transcendental Phenomenology, Edmund Husserl brings light to the indeterminacy that he finds in every act of intentionality. He does this through

Genetic Phenomenology, Indeterminacy of Racism  169 the idea of horizon. Gail Weiss has elaborated upon Husserl’s use of horizons in our everyday experiences and their relationship to indeterminacy.1 She focuses on indeterminacy within everyday constitutive acts with respect primarily to temporal horizons. She highlights passages such as the following where Husserl suggests that intentional acts, necessarily imply an infinite horizon of inactive validities which function with them in flowing mobility. The manifold acquisitions of earlier active life are not dead sediments; even the background (for example that of the perceptual field), of which we are always concurrently conscious but which is momentarily irrelevant and remains completely unnoticed, still functions according to its implicit validities.2 In other words, constitution takes place within a horizon of sedimented validities that we only make apparent to ourselves through a phenomenological peeling away of layers of previous experiences. Weiss claims that “indeterminacy can be viewed to a large extent as arising out of the temporality of human existence.”3 One’s own history comprises the temporally prior experiences that color constitutional acts in the present in ways of which we are frequently unaware. Let us consider an example. Even now as I write this chapter or as you read this chapter, your focus is upon the words and meaning conveyed therein. But we are both likely seated in a room with ambient sound and light. These elements of our embodied experiences are most likely not the focus of our constitution until they are brought to our attention, as has happened in the bringing up of an example. At the same time, there are other layers of experience coming to bear, as well. You expect there to be a structure to this argument that will ideally make sense. You draw upon your own previous encounters with Husserlian phenomenology to understand the references. You also probably could bring some of that previous experience more or less to light for yourself if you consider the exposure to Husserlian phenomenology that you have had, what the context of that exposure may have been, whether there might have been any particular differences in how Husserl may have been framed for you as opposed to how he was framed for this author. That may have something to do with where each of us was educated, who our teachers were, etc. The “etc.” here indicates that this horizon of validities is infinite. It entails every aspect of the experience. Even the act of reading is stylized by prior experience. Again, Husserl explains, and as Weiss has quoted, every straightforwardly performed validity in natural world-life always presupposes validities extending back, immediately or mediately, into a necessary subsoil of obscure but occasionally available reactivatable validities, all of which together, including the present acts, make up a single indivisible interrelated complex of life.4

170  Janet Donohoe Tracing back these validities is the genetic phenomenological practice and it allows us to investigate that interrelated complex of life in such a way that we can reactivate validities or renew them while at the same time being able to critique those same validities. For Husserl, all acts of intentionality have a horizonal nature that can become apparent through a questioning back (Rückfrage) into the sedimented layers of our everyday experience. The aim of Rückfrage is to make apparent the aspects of any constitution that have previously been so much a part of the horizons of the constitution that we have failed to notice them or pay heed to their functioning. By asking back after them, we can more attentively reactivate these elements and recognize the ways in which we have perhaps assumed aspects of constitution. For instance, perhaps in the reading of a phenomenological paper, I discover that my own understanding of a particular concept entails an emphasis that is not found within the author’s approach. I may determine through my own asking back that I picked up that emphasis from a professor or mentor of mine. It is now available to me for examination, and I may either reaffirm that emphasis as being appropriate, or I may reject that emphasis based upon the article I am currently reading. The current article has perhaps given me a perspective from which to critique the aspect of the concept that I previously thought was simply inherent in the concept itself, but now realize was a kind of shading of the concept based upon my exposure to it through a mentor. While Weiss recognizes three different types of horizons, she focuses only on two, which she terms the “internal” horizon and the “external” horizon. The internal horizon refers to the “multiplicity of possible perceptions I can have of a given thing,”5 while the external horizon “points toward the world which serves as the continuous horizon for all of my actual and possible experiences.”6 Weiss acknowledges the third type of horizon in a footnote as “the world horizon which situates the perceptual field.”7 For the sake of clarity, I suggest that the external horizon is homeworld or alienworld. In other words, it is the immediate perceptual field that is sedimented in the constitution of a thing. The third horizon I would identify as the pre-given lifeworld. Associating horizons with these levels of homeworld, alienworld, and lifeworld means that each is also associated with some indeterminacy. Suggesting that there is indeterminacy even at the level of the pre-given lifeworld will have major repercussions as we will see below. We must first explicate the different types of horizons to see in what way each is associated with indeterminacy. The first type of horizon refers to the many ways in which any object of perception is given to me. The computer in front of me is given to me as a computer, even though what my senses actually tell me about it are that there is a cold, relatively flat piece with a vertical screen emanating light. My brute sense impressions do not give me computer; they give me various visual and touch sensations which do not equate to computer. But what

Genetic Phenomenology, Indeterminacy of Racism  171 I experience is a computer that is given to me in its wholeness. I do not experience brute sensations. The internal horizons of this experience entail the parts of the computer that are not immediately sensorily present to me. They enable me to experience the object as computer and not as dissociated sensory impressions. My focus can shift with those horizons to various aspects of the thing that contribute to the givenness of it as computer. Thus, the horizon of the sensations gives me a sense of the other aspects of the computer that are not immediately sensorily perceived, but that are given to me in the givenness of the computer. The second type of horizon, or the “external” horizon, refers to the perceptual field where we can focus on this or that aspect of a thing’s immediate environment. As I sit at my computer, the perceptual field of my constitution of the computer includes the room within which I sit and the glass of water near me on the desk, as well as the music playing in the background, and the smell of dinner simmering on the stove. Beyond the computer screen and my intent to get some thoughts typed out, however, the rest is more or less indeterminate. The things upon the desk, the ambient sounds and smells are not the focus of my constitution and are, thus, horizonal aspects of my constitution of the computer screen. They are indeterminate insofar as they are not my focus and I am only vaguely aware of them, if not completely oblivious to them. The glass of water may become momentarily more determinate as I reach for it to quench my thirst, but I may also reach for it without paying a whole lot of attention to it. I reach for it out of habit and without moving my gaze away from what I am typing on the computer. Until I accidentally knock the glass over and must quickly respond, it is merely indeterminately within the horizon of my computer screen. When the glass becomes my panicked focus, it becomes more determinate and the computer screen retreats into indeterminacy as I am distracted by the glass. If the glass were to actually tip over and break, it would become even more determinate as I begin to focus very specifically upon its shards of glass in order to be vigilant against getting cut. As my focus shifts from one object to another in the room, however, the objects that are not my focus do not disappear. They become differently and more indeterminately constituted as horizon. As one object becomes my focus, others lose that focus becoming more indeterminate. Indeterminacy, then, shifts as focus shifts such that it is always present in any constitutive act. It is only against this backdrop of indeterminacy that anything is determinate at all. At the same time, we are never without horizons no matter how much of a laser focus we might have upon any particular thing. The second layer of horizon brings to bear different aspects of the constitutional act. It aligns with Husserl’s discussion of homeworld and alienworld while the third layer of horizons aligns with lifeworld. In the following section, I will detail these notions of horizon with respect to the Husserlian conception of homeworld and alienworld and how they contribute to a

172  Janet Donohoe Husserlian idea of indeterminacy. We will also consider how the third level of horizon, lifeworld, contributes to this view and how we can respond to the important ramifications of indeterminacy. But first it is important to clarify what we mean by givenness and pregivenness. Primarily, we’ve been speaking of givenness so far. Givenness refers to the objects of experience that are not reducible to the particular aspect of the thing that might be the perspective of our experience at the moment. In other words, that which is given is always within a horizon of possible aspects. As described, the computer is given to me as computer even though I do not see the back side of it or the interior of it. The fact that I do not see these other aspects does not reduce the computer to the bare empirical sensory perception. The computer is given to me in its completeness as a computer. Pre-givenness, on the other hand, helps us to understand an even deeper and more indeterminate level of constitution. It is the ground upon which all givenness rests. The pre-given is always already there for us in our constitution of anything whatsoever. It is what is constantly valid and upon which any experience, actual or possible, depends. The pre-given world is that of which we are always already aware before we consciously turn towards it in analysis or reflection. It is lifeworld. Although horizons cannot be made objects of experience, they are entailed in every experience, for all experience is horizonal. We do not ever experience objects without background or surrounding world of alienworld or homeworld. The horizon that is lifeworld is the very condition of any objects of world as being given. Lifeworld is pre-given rather than given. As pre-given, the lifeworld does not necessarily have its own peculiar sense nor can it be made an object of sense. It is rather the condition of the possibility of any particular given homeworld or alienworld. This means that it is the indeterminate horizon of all experience.

Homeworld/Alienworld Husserl is interested in the ideas of homeworld, alienworld, and lifeworld in an effort to come to grips with the notion of an objective world. He recognizes that traditions and history have an impact on our experiences of our world in such a way that might cause difficulties for a notion of an objective world. To grapple with that difficulty, he asserts that homeworld and alienworld are parts to lifeworld’s whole, but not in the sense that the sum of the parts compose the whole. Rather, one understands the parts because of the understanding of the whole, and vice versa. Lifeworld can only be grasped through homeworld, and homeworld is grounded in and made possible by lifeworld. Lifeworld, in other words, is the horizon for homeworld, and is that which is pre-given in the givenness of a homeworld. If we return to the example of the computer at which I sit and its horizons of desk and water glass and room, etc., we grasp that the familiar

Genetic Phenomenology, Indeterminacy of Racism  173 ways in which I interact with these things has to do with very personal ways in which my movements through the world have been habituated. The horizons include sedimented, cultural manifestations of traditions in which I have been raised and which I have developed through prior experiences. My typing on the keys of the computer is so habitual that I do not pay any attention to the keys themselves as much as I pay attention to how I want to formulate a sentence or a thought. The ambient sounds and temperature are horizonal insofar as they are not my focus, but also insofar as they reflect a certain tradition and sedimented homeworld. Were the sounds of the horizons of my experience suddenly to change, for instance if I were to hear an elephant or perhaps a hot-air balloon outside my window, they would draw my attention as not being “normal” and would intrude into my homeworld with something alien. Clearly, these sounds are not alien to everyone. They are alien to me and my homeworld. I could have a completely similar experience of typing on my computer at a desk with my water glass in Albuquerque, New Mexico, where the sound of a hot-air balloon might not be alien since it is, after all, the hub for the International Hot-Air Balloon Races each year. The fact that such a sound is alien in my homeworld helps us to see that horizons are culturally sedimented and require us to peel away layers of that sedimentation to bring to determinacy things previously indeterminate. And while the environment in which the sound of an elephant outside my window might be radically different from my own, it too would share certain, perhaps deeper, levels of horizon, lifeworld horizon. There are levels of similarity that have to do with historical position, but also have to do with ways in which the lifeworld acts upon us. Such similarities allow me to recognize an alienworld as a world with its own horizons. Albuquerque, New Mexico, or a place where elephants are common are both places that are culturally and physically different from the study where I compose philosophical articles, but they are all part of a lifeworld that makes both those worlds and my homeworld possible. The horizon of lifeworld that makes the homeworld and alienworlds possible is not an objective, unified, unchanging world, however. It is less determinate than that. The lifeworld is pre-given in the sense that it functions as the indeterminate horizon, or the background of any experience or encounter with homeworld or alienworld. Homeworld is the indeterminate horizon of each of my experiences and includes my own sedimented experiences as well as cultural meanings that have been passed along to me through the world in ways that I do not necessarily even recognize. In my everyday experience I do not even consider that the ambient sounds are “normal” or that the sound of an elephant would not be normal. I take completely for granted the sounds that are always the horizon of my experiences in the study. Until I shift my focus and really listen to those sounds, they are indeterminate horizonal sounds. The indeterminacy of such horizons arises in large part because we each take on certain expectations about our world and our experiences. We

174  Janet Donohoe develop those expectations from the moment that we begin to have experiences. Those expectations become so ingrained as to be accepted as simply the way the world is in itself. These assumptions allow for much of the world to remain indeterminate. We expect the sounds of cars or children in the school playground, for instance, such that we no longer even notice them as sounds. Experiences vary in the degree to which they are more clearly cultural and relative to one’s own culture or even one’s own family or home life situation than others. When something unexpected happens, or when one experiences something or someplace that is not familiar, it is experienced as an alienworld and can shed light on that indeterminate homeworld horizon helping to make it more determinate. Consider focusing on the pages of a book in an everyday kind of way. One constitutes the book as something to be read, perhaps marked upon and digested, so to speak. This activity is one that others would look upon without necessarily even acknowledging it as in any way significant. It is an ordinary activity in everyday life. It might mark one out as being an educated individual or perhaps as being of a particular class, but it is within a homeworld in which it can have such significance. In the reading of the book, there is much that is given. The constitution of the book has horizons that are completely taken for granted as part and parcel of the constitution without themselves being the focus of the constitutive act or even being noted as part of the constitutive act. The continuity of the book is given. We presume that the pages of the book are continuous and will be written in the same language throughout. We presume that the pages follow in order one from the other. But even beyond that, we constitute the book as something to be purchased or borrowed from a library. These constitutive expectations hold within them layers of sedimentation of our personal histories, but also homeworldly significant sedimentations as described above. We may only recognize these as homeworld elements when we are confronted with the alienworld of another culture, perhaps a culture where a book holds a completely different cultural designation and indicates a completely different class of person or activity. Even the activity itself is different. While my eyes are thoroughly trained to read left to right and top to bottom, this is given in the constitution of the book. It’s so deeply a part of what is given in the constitution of a book as to be thoroughly overlooked. But, if I were handed an Arabic text, I would find myself completely flummoxed. It would be difficult to constitute the book as a book I could read, and its alienness might cause me to realize my own assumptions about how a book is to be read and what comprises a book. It is difficult for me to think through this difference as regards the book until I have engaged in a genetic phenomenological approach. I begin to peel away layers of my experiences of books, discovering assumptions that I have about books, and what constitutes a book, and even where those assumptions come from and how they are connected to my homeworld. Examining my assumptions about a matter is frequently motivated through

Genetic Phenomenology, Indeterminacy of Racism  175 an encounter with an alienworld. If one never encounters a book that reads in any way other than from left to right and top to bottom, then one might be inclined to think that to be a book entails such a matter of reading. Thus, reading left to right, top to bottom is constituted in the givenness of the book itself. However, once one begins the process of peeling away the sedimented layers, one can recognize that what is given is not an absolutely determinate way of being a book. Being a book entails the indeterminacy of unacknowledged levels of givenness. These characteristics can be brought more or less to the fore through the process of renewal and critique, but only once they are brought to light through the genetic phenomenological method. What also becomes revealed in an encounter with something unexpected is the way in which homeworld and alienworld are both made possible because of the more fundamental lifeworld. To recognize an alienworld trait as alien, it must be within parameters that make some sense. Something completely outside of the realm of sense would not be available for experience at all. The lifeworld is the horizonal level that allows for the possibility of experience at all. It is fundamental to our being of this earth. Homeworlds of vastly different kinds share lifeworld. The earthworm homeworld, the plankton homeworld, and the human homeworld overlap in many ways, but are vastly different in other ways. They share lifeworld as ground and horizon, making each of them possible and making each of them unique to this earth. Worm, plankton, human, we are all earthlings. We are earthlings in our manner of constitution which distinguishes us from beings not of this Earth. It is telling that in our own explorations of space, we eagerly search for earth-like planets. Yet, were we to find an Earth-like planet, it would not be constituted by its natives in an Earth-like way. Likewise, if we were to transport ourselves to Mars, we would not become Martians. Rather, as Earthlings, we would constitute Mars in an Earth-like way. In many respects, we cannot even identify what constitution in an Earthlike way is because we have no alien world with which to contrast our own lifeworld. This means that much of our lifeworld necessarily remains indeterminate because we remain blinded to its aspects that are so thoroughly taken for granted as to be hidden from us. These are the layers of pre-given lifeworld. It is important to keep in mind that this pre-given lifeworld is not an objective world. It is one that is open to the dynamics of historical change. This means that the contemporary pre-given lifeworld entails presuppositions that it may not have entailed in previous epochs. For instance, the existence of fossil fuels is an assumed aspect of our everyday lifeworld that is taken up differently by different homeworlds, but which was not an aspect of the medieval lifeworld, for instance. The concomitant destruction of our environment as a result of the burning of fossil fuels is an aspect of the Western homeworld that has perhaps allowed us to come to grips with certain other aspects of our lifeworld that we have thus far taken completely for granted. Without the impending destruction of our environment, we

176  Janet Donohoe might not be motivated to engage with the process of critique of our traditions of consumption that support our continued need for fossil fuels. The layers of givenness and pre-givenness we have been exploring indicate layers of indeterminacy. These function at both the level of homeworld and alienworld, but also at the level of lifeworld. Interestingly, if we consider the examples of the computer and the book, we come to recognize that what is at first thought to be indeterminate can be made more determinate through the focus of our attention as well as through other encounters. But, it is also apparent that indeterminacy can never be made completely determinate. While we can plumb the depths of our own experiences, and focus on more and more aspects of a thing, its background, its edges, its other sides so to speak, the task is infinite meaning that indeterminacy is a fundamental element of any constitutive act and, therefore, of our world. Pre-given lifeworld is the horizon of all experience. Every experience entails presuppositions about the way the world is without investigation into the indeterminacy that is taken up by each of us. We are unaware of our style of constitution and lifeworld remains unthematized precisely because of its horizonal nature. And yet, because lifeworld is the very condition of a cultural homeworld or alienworld it is the ground of every experience.

Ethical Ramifications/Renewal and Critique How can we or how should we respond to this prevalence of indeterminacy? On the one hand, we might be inclined to throw our hands in the air and lament the infinite task of attempting to make determinate that which is indeterminate. We could wallow in our despair and decide that indeterminacy is good enough, that we have no reason to strive for determinacy. On the other hand, we could double-down on our efforts to bring to determinacy that which is indeterminate. This latter approach is what Husserl would advocate, I suggest. The indeterminacy that we accept within our lifeworld is not always benign. The examples I have used thus far in describing the homeworld/alienworld/lifeworld triad are benign making it perhaps not quite so clear why there would be ethical undertones to this discussion. Let us make the ethical ramifications more explicit. It is not an uncommon occurrence for blacks in the United States to be pulled over or surveilled if they are seen in a predominantly white suburban neighborhood. Whites living in such neighborhoods are habituated to seeing other white people, and primarily only white people. So, if white people in such a neighborhood see a black person they may be surprised by this experience and it may challenge their presuppositions about their own neighborhood. Their surprise, in such a case, is frequently accompanied by a feeling of anxiety, due to deeply sedimented cultural notions of black people as more inclined to be criminals or to be violent or dirty. These sedimented notions have been passed along for generations among whites

Genetic Phenomenology, Indeterminacy of Racism  177 and have been cause for keeping black people from being comfortable in a white-organized world. Sara Ahmed draws upon the work of Frantz Fanon to describe it thus: As Fanon’s work shows, after all, bodies are shaped by histories of colonialism, which makes the world “white,” a world that is inherited, or which is already given before the point of an individual’s arrival. This is the familiar world, the world of whiteness, as a world we know implicitly. . . . Bodies remember such histories, even when we forget them. Such histories, we might say, surface on the body, or even shape how bodies surface. Race then does become a social as well as bodily given, or what we receive from others as an inheritance of this history.8 While Ahmed’s emphasis is on the bodies of whites and blacks, she also notes that the world is familiar for whites. Such familiarity is about that world being the white homeworld in which the black person is alien. She recognizes that these attitudes are given in the world itself. What she describes is what we might call systemic racism. Further, Ahmed writes, “To think of this implicit knowledge as inherited is to think about how we inherit a relation to place and to placement: at home, things are not done a certain way, but the domestic ‘puts things’ in their place. Whiteness is inherited through the very placement of things.”9 The world of things that is given involves the horizons of homeworld that are inherited and which we can reveal through the genetic phenomenological method in order to leave them open to the possibility of critique. Ahmed underscores the way that whiteness functions as a horizon for whites in a different way than it does for non-whites. She claims that: “Spaces are orientated ‘around’ whiteness, insofar as whiteness is not seen. We do not face whiteness; it ‘trails behind’ bodies, as what is assumed to be given. The effect of this ‘around whiteness’ is the institutionalization of a certain ‘likeness,’ which makes non-white bodies feel uncomfortable, exposed, visible, different, when they take up this space.”10 Ahmed is describing a homeworld horizon for both whites and non-whites. To call it a homeworld might imply a certain level of comfort, but that is not exactly what Husserl means. Homeworld indicates a level of “normality” that does not imply good or bad. For non-whites, this homeworld entails a level of discomfort that whites do not face. It is no less sedimented, however. Non-whites have been living in this uncomfortable homeworld for generations, just as whites have been living with their privilege and comfort for generations such that for many it is viewed as simply the way the world is. Husserl is not at all suggesting that homeworld should remain unchallenged. Just because homeworld might be sedimented and fulfill expectations does not make homeworld good or right. We might wonder how anyone has resources for constituting homeworld any differently. For Husserl, this involves the responsibility for renewal and critique.

178  Janet Donohoe There are many levels of racism going on here. Privileged white people have the luxury of being oblivious to their whiteness, which is why this is a horizon of their experience. Should they encounter the kind of surprising experience just described, they might come face-to-face with their own prejudice by examining why they are surprised and thinking more deeply about their expectations and why they have those expectations. This is the process of critique that Husserl is advocating. When confronted with what is alien, one should not simply reaffirm one’s own homeworld. The alien makes homeworld apparent to one in such a way that it is open for critique. It is also open for renewal, but for Husserl renewal is never separate from critique. It may be the case that one chooses to renew the tradition one has inherited, but if one does so without critique then one is being irresponsible. The process of renewal as described by Husserl in The Crisis of European Sciences is “to make vital again, in its concealed historical meaning, the sedimented conceptual system which, as taken for granted, serves as the ground of [the individual’s] private nonhistorical work.”11 Note that Husserl recognizes that the historical meaning is concealed until one revitalizes it in the process of renewal. Husserl is very clear that one ought not stop with renewal, however. Instead, the goal is to make present for ourselves the tradition that underpins a position by connecting it to the “chain of thinkers, the social interrelation of their thinking, [and] the community of their thought.” Further, Husserl demands that we “carry out a responsible critique, a peculiar sort of critique which has its ground in these historical, personal projects, partial fulfillments, and exchanges of criticism rather than in what is privately take for granted by the present philosopher.”12 For Husserl, critique is inseparable from the process of taking up one’s homeworld and the demand for critique is an ethical responsibility that leads to a better humanity. This discussion of renewal and critique reveals even more forcefully the importance of the genetic phenomenological method. It is in peeling back the layers of sedimented tradition that are normally functioning in their indeterminacy, that we are in a position to evaluate and critique, making positions and traditions more and more determinate. We might wonder what the impetus would be for any person to choose to change based upon a critique of their own tradition rather than to simply renew the tradition. For Husserl, the responsibility for critique entails responsibility for the community. The community does not function separately from its members, meaning that the self-responsibility of the individual is the foundation for the self-responsibility of the community. The aims of the individual entail the aims of the Other and contribute to what Husserl calls the “higher-order we.”13 Since this community is of a higher order, it must be grounded upon the freely acting individuals that compose it. This means that the individuals cannot be absorbed into the community, but instead absorb the communal goals into their own. This relationship

Genetic Phenomenology, Indeterminacy of Racism  179 mitigates against a communal domination over the individual and against an imperialist unity of the community. Husserl argues that imperialist unity of will is irrational, since the sense of a hierarchy of power that it suggests places an individual or group in a position of power over others, thereby eliminating the true rationality of free position-taking. But Husserl recognizes that he has no way of ensuring that the individuals composing the community will take up the responsibility for renewal and critique. He does believe, however, that the community of the higher-order we is antitotalitarian and antiauthoritarian but does not admit of a system of external ethical norms. For him, ethics does not mean normative morality. Rather, it depends upon the self-critique that leads to personal responsibility and ultimately relies upon a position of ethical love. If one takes on the responsibility for critique, one does so not for one’s sake alone, but also for the sake of the community of others.14 In the situation of the person who is immediately suspicious of a black man encountered in her “white” neighborhood, she can be brought to focus on the indeterminate aspects of her encounter by the process of critique. She should ask herself why it should be alien for a black person to be in her neighborhood and while she may recognize that it is not “normal,” this does not necessitate that it be threatening, as opposed to simply curious. She ideally would be compelled to critique the tradition that leads her to believe without any evidence that the black person’s presence in her neighborhood is something to be suspicious of. Without any evidence of suspicious behavior she could possibly be challenged to rethink her own racist response that is more than likely something that she inherited from previous generations. Her inheritance of this position, however, does not absolve her of the responsibility to critique the position before renewing it. We would hope that she would be inclined to work towards not renewing such a tradition, but attempting to replace that tradition with a more just tradition. The ethical responsibility that Husserl endorses, then, is one that is internal rather than external. Given the indeterminacy inherent in that which is pre-given, the responsibility for renewal and critique is infinite. Even when we suggest the possibility of a more just tradition, it cannot be the tradition with which we settle. We are called to renew and critique again and again such that universal norms cannot simply be accepted as norms any more than cultural norms can be accepted. We know that the scenario of a person of color perceived to be in the “wrong” neighborhood is an all too common occurrence and yet, the outcome rarely seems so positive as inclining a racist person to rethink her own racism. The rarity of such an outcome underscores the difficulty of making determinate that which is indeterminate, or bringing to the fore those deepseated ways in which we constitute the world that are horizonal and difficult to bring into focus. This analysis perhaps offers us a mechanism to be able to discuss the process that is necessary in moving beyond white privilege and hidden racism.

180  Janet Donohoe

Conclusion So, if what is pre-given is indeterminate, this means that at the core of every constitutional act is indeterminacy. Why should we find this so astonishing? We cannot, no matter the degree of critique we take on, make that which is indeterminate completely determinate. Does this negate our struggle against indeterminacy? Not entirely. Instead, we recognize that the role of renewal and critique helps us to bring into focus more and more of the indeterminate while recognizing at the same time that there is always that dark core of indeterminacy that shades constitution. Rather than indeterminacy being something we might use to absolve ourselves of our responsibility by claiming that we cannot possibly be responsible for ways of constitution that we have inherited without being thoroughly conscious of having done so, Husserl’s call to renewal and critique requires of us a more vigilant position. We must not blindly adhere to positions that we have inherited without first attempting to ask back into the ground of those positions. We are called to peel away the layers of constitution, shine a light on the horizons of our acts to make more and more clear to ourselves how our constitution is shaded by indeterminate horizons. We have an obligation to rigorously critique those horizons making every effort to bring them to determinacy all the while knowing that our attempt can never be complete.

Notes 1 See Weiss (1995). 2 Husserl (1970, p. 149). 3 Weiss (1995, p. 44). 4 Husserl (1970, p. 149). 5 Weiss (1995, p. 44). 6 Ibid. 7 ibid., p. 51fn2. 8 Ahmed (2007, pp. 153–154). 9 Ibid., p. 155. 10 Ahmed (2007, p. 157). 11 Husserl (1970, p. 71). 12 Ibid., pp. 71–72. 13 See Donohoe (2016, especially pp. 135–145). 14 Husserl’s notion of ethical love is not fully developed, but is linked to his explanation of each person’s absolute ought which is a personal commitment to values. For example, a mother’s absolute ought might be providing for her children. They have absolute value for her. But she might also have an absolute ought in the realm of being a scholar. These oughts are absolute insofar as they define her ethical commitments and cannot therefore be placed into some kind of normative hierarchy. For more on this, see Donohoe (2016, especially pp. 129–135).

Works Cited Ahmed, Sara. 2007. “A Phenomenology of Whiteness.” Feminist Theory 8(2): 149–168.

Genetic Phenomenology, Indeterminacy of Racism  181 Donohoe, Janet. 2016. Husserl on Ethics and Intersubjectivity: From Static to Genetic Phenomenology, 2nd ed. Toronto: Toronto University Press. Husserl, Edmund. 1970. The Crisis of European Sciences and Transcendental Phenomenology, translated by David Carr. Evanston, IL: Northwestern University Press. Weiss, Gail. 1995. “Ambiguity, Absurdity, and Reversibility: Responses to Indeterminacy.” Journal of the British Society for Phenomenology 26(1): 43–51.

8 Indeterminacy as Key to a Phenomenological Reinterpretation of Aristotle’s Intellectual Virtues Robert H. Scott In book 1, section 7 of Nicomachean Ethics, Aristotle presents what has become known as “the function argument,” and he frames it as a way to better explain the nature of happiness. The argument starts with the premise that a thing can attain its highest end by performing its function well, and it proceeds by way of an analogy: just as the eye attains its highest end of seeing well by performing its function well (e.g., seeing), so a human being can attain its highest end (happiness, or eudaimonia) by performing the human function well, whatever the latter may be. Aristotle then sorts through possible candidates for the specific character of the human function by distinguishing distinctly human activity from that of plants and nonhuman animals, finally arguing that the human function must be to act according to a rational principle, which he equates with virtue (or human excellence), because rational activity is what separates humans from plants and other animals.1 Returning to the analogy, Aristotle goes on to argue that just as other things attain their highest end by performing their function well, human beings must attain their highest end (happiness) by excelling in acting according to a rational principle, and this Aristotelian insight remains an important point of departure for virtue ethics theory up to the present. On the other hand, a departure from Aristotle’s account of the theoretical intellectual virtues (in book 6 of Nicomachean Ethics) is called for, I contend, due to its narrow teleological orientation toward attaining completely determinate knowledge and its correlative privileging of theoretical wisdom over practical wisdom. In this chapter, I argue, through a phenomenological disclosure of horizons of indeterminacy surrounding all objects of understanding in the phenomenological field of inquiry (e.g., the world horizon), that the highest form of reason (philosophic wisdom) must include recognition of horizons of indeterminacy and that these horizons carry ethico-practical significance that is integral to the meaning of knowing. In view of the disclosure of the ethico-practical significance of horizons of indeterminacy, I further argue that a phenomenological reinterpretation of the intellectual virtues serves to clarify, in a way that Aristotle did not, the necessary, integral interdependence of theoretical and practical reason.

Indeterminacy as Key to Reinterpretation  183 In what follows, I first present an abbreviated transcendental phenomenological theory of reason which I then apply in reinterpreting and retrieving the Aristotelian theoretical intellectual virtues of intuitive reason (noûs), scientific knowledge (epistéme), and philosophic wisdom (sophía).2 An important advantage of approaching Aristotle’s account of the intellectual virtues through the lens of transcendental phenomenology is that the latter is unburdened by a problematic substance ontology, integral to Aristotle’s account, and this unburdening opens the way to a fluidly unified theory of theoretical and practical reason. The key phenomenological insights that lead to a more unified conception of the interdependence of theoretical and practical reason include the disclosure of the “world horizon” as the field of transcendental phenomenological inquiry and, subsequently, of horizons of indeterminacy at the limits of a clear understanding of any object.3 The disclosure of horizons of indeterminacy surrounding objects leads to the further disclosure that the meanings of objects necessarily interconnect, through their horizons of indeterminacy, with other horizons of significance in the world horizon, and they do so in ways that carry important ethicopractical implications. By reinterpreting the three theoretical Aristotelian intellectual virtues in view of horizons of indeterminacy, I conclude that a transcendental phenomenological theory of reason contributes to clarifying the necessary, non-hierarchical and integral interdependence of theoretical and practical reason. In book 6 of Nicomachean Ethics, Aristotle introduces and explains the five intellectual virtues as constitutive of the overall excellence (e.g., virtue) of the rational part of the soul, and, in doing so, he further delineates a strong distinction between the activity of theoretical and practical reason, based on the teleological orientation of the soul towards pure contemplation. The five intellectual virtues he identifies, as is well known, are art (téchne), practical wisdom (phrónesis), intuitive reason (noûs), scientific knowledge (epistéme), and philosophic wisdom (sophía), the first two being virtues of practical reason and the latter three being virtues of theoretical reason.4 Aristotle’s basis for drawing a strong distinction between practical reason and theoretical reason rests on the relation between the end of pure contemplation and the difference between the kinds of objects each kind of reason treats: practical reason is concerned with variable objects, and the two intellectual virtues that enact this form of reasoning are practical wisdom, which guides moral virtue by providing the rule of the mean for action, and art which is concerned with making and building.5 Theoretical reason, by contrast, is concerned with invariable objects, namely first principles and what follows from them.6 It is important to keep in mind the strong distinction Aristotle draws between practical and theoretical reason, for it leads him to later draw a value-hierarchy between the two (in book 10) that prioritizes theoretical reason over practical reason in the pursuit of the highest form of happiness. In book 10 of Nicomachean Ethics, Aristotle indicates that the highest level

184  Robert H. Scott of happiness is achievable through the virtue of philosophical wisdom, and he supports this claim by noting that the latter is the highest virtue of the highest kind of reasoning (theoretical reason). He further emphasizes the hierarchical distinction between theoretical and practical reason by saying that the happiness achievable by the other kind of reason (practical reason) is of “a secondary degree.”7 By reinterpreting the intellectual virtues through the lens of a transcendental phenomenological theory of reason, I call into question the strong distinction and hierarchy Aristotle draws between theoretical and practical reason, especially insofar as Aristotle allows for theoretical reason to detach from practical reason in staking a claim to being the highest form of reasoning. Further, I argue that a key insight of transcendental phenomenological reason, taken as a normative form of rationality, consists in its disclosure of perennial horizons of indeterminacy at the limits of clear understanding that bind objects of theoretical reason with other horizons of meaning in the world horizon, the latter of which includes binding theoretical objects with important ethico-practical demands. The grounds on which Aristotle establishes a value-hierarchy of theoretical over practical reasoning can be traced back to his metaphysics and, in particular, to his theory of the immaterial primary cause (the Unmoved Mover) according to which higher value is assigned to activities and things that approximate most closely to the activity and substance of the first cause.8 For Aristotle, the activity of theoretical reason approximates more closely to the activity of the Unmoved Mover than does practical reason because the objects of theoretical reason are invariable, immaterial truths while the objects of practical reason are variable and intermixed with matter.9 Another way to say this is that the objects of theoretical reason, in Aristotle’s account, avail themselves of complete determinacy; such objects include first principles, mathematical objects, and empirical objects of scientific study. By contrast, the objects of practical reason, including products of making and applications of the moral rule of the mean, are bound up with matter and worldly concerns and, as variable, include some degree of indeterminacy which, for Aristotle, lowers the value of practical knowing in relation to theoretical knowing.10 Conversely, for Aristotle, because the activity of theoretical reason approximates more closely to the pure, immaterial activity of the Unmoved Mover than do the activities of practical reason, he distinguishes and ranks theoretical reason above practical reason.11 Reflecting on Aristotle’s account of theoretical and practical reason, we may question why invariable objects should be considered superior in value to variable objects and, correlatively, why theoretical reason should be considered superior to and separable from practical reason. As indicated above, the ranking of theoretical reason above practical reason stems, for Aristotle, from the proximity between the immaterial, unchanging character and activity of the Unmoved Mover and the theoretical rational activity of grasping invariable objects. Importantly, from a transcendental phenomenological perspective, under the phenomenological reduction there is no need to

Indeterminacy as Key to Reinterpretation  185 suppose an immaterial Unmoved Mover as primary cause,12 and, by setting aside (bracketing out) this supposition, the motivation for detaching theoretical activity from practical reason and ranking the former (along with its objects) above the latter (and its objects) dissolves. By unburdening itself of a problematic substance ontology (or never introducing the idea or setting it aside), a transcendental phenomenological approach to understanding has no motivation to posit the attainment of fully determinate knowledge as its primary goal. Having set aside the assumption of the Unmoved Mover, a transcendental phenomenological theory of reason calls into question both the possibility and desirability of attaining entirely determinate theoretical understanding as the primary goal or telos of reason. To the contrary, a transcendental phenomenological theory of reason serves to clarify that a completely determinate understanding of objects is neither possible nor desirable as a goal of reason, and this is the case because, while mathematical objects and other theoretical objects may be fully determinable from a limited, epistemological perspective internal to themselves, all objects within the reduced field of transcendental phenomenological inquiry maintain an external horizon of indeterminacy which links their meaning with other horizons of meaning within the world horizon.13 The overlapping of horizons of meaning in the world horizon, generated by horizons of indeterminacy surrounding objects, comes to be understood, through transcendental phenomenological inquiry, as integral to a rational understanding of any object. It is further important to note that the phenomenological disclosure of horizons of indeterminacy surrounding objects along with the intertwining of the meanings of objects with other horizons of significance leads to the further disclosure of ethicopractical demands that tie practical and theoretical reason together. Before explaining how transcendental phenomenology discloses horizons of indeterminacy at the limits of clear understanding and how such horizons point to ethico-practical demands and the integral interdependence and unity of theoretical and practical reason, I briefly turn now to explain further how Aristotle’s metaphysics leads him to draw an excessively sharp division between theoretical and practical reason and to privilege the former over the latter. The hierarchical division of the intellectual virtues implicit in Aristotle’s account can be traced back to his teleological metaphysical theory, according to which the immaterial Unmoved Mover stands as the final cause of all things. In view of this onto-theology, we can make sense of Aristotle’s subordination of practical reason to theoretical reason in that, due to the material character of objects of practical reason (what he calls “variable objects”), the latter bear less affinity to the immaterial Unmoved Mover than do objects of theoretical reason which he characterizes as invariable. Similarly, he considers the activity of theoretical reason, in grasping invariable objects, to have the capacity to be immaterial and more god-like than the activity of practical reason, which is bound to matter (in making) and

186  Robert H. Scott worldly concerns (in applying the rule of the mean). What, we may ask, becomes of the strong distinction Aristotle draws between theoretical and practical reason if we bracket out, as the phenomenological reduction does, the idea of the Unmoved Mover as final cause and the onto-theological teleological orientation that comes with it? In the following section, I argue that horizons of indeterminacy provide a key to a phenomenological reinterpretation of the intellectual virtues by showing that a thorough rational understanding of any object involves recognizing horizons of indeterminacy at the limits of clear understanding and effectively responding to ethico-practical demands that arise from these horizons. After distinguishing a transcendental phenomenological understanding of reason from Aristotle’s approach, I consider how the character and ordering of the three primary Aristotelian theoretical intellectual virtues (intuitive reason, scientific knowledge, and philosophic wisdom) would change when approached through the lens of a transcendental phenomenological theory of reason. As will be seen, not only does a transcendental phenomenological retrieval of the theoretical intellectual virtues disclose the manner in which all inquiry and knowledge of objects is embedded in a manifold context of overlapping, interconnecting (and meaning enriching) horizons of indeterminacy in the world horizon, it further serves to disclose ethico-practical demands and the necessary interdependence and unity of theoretical and practical reason.

A Transcendental Phenomenological Theory of Reason and the Significance of Horizons of Indeterminacy In The Crisis of European Sciences and Transcendental Phenomenology,14 Husserl recalls and aligns his project with the rationalist tradition that goes back to the ancient Greeks. He writes, It is reason which ultimately gives meaning to everything that is thought to be, all things, values and ends—their meaning understood as their normative relatedness to what, since the beginnings of philosophy, is meant by the word ‘truth’—truth in itself—and correlatively the term ‘what is’—ὄντωσ ὄν.15 In this passage, Husserl embraces the classical and modern rationalist idea that reason is the ultimate source for a normative understanding of truth and being; however, he goes on to argue in Crisis that transcendental phenomenology transforms the rationalist project by means of the transcendental phenomenological reduction which brackets out the naïve treatment of both objects and the ego as separate entities and clarifies the field of inquiry as the manifold sphere of evidence given to intuition in what he describes as the world horizon.16 Under the transcendental reduction, the field of study becomes the manifold of phenomena as they appear in the world horizon,

Indeterminacy as Key to Reinterpretation  187 and there is a re-focusing of attention and analysis away from objects as existing in the external world towards “the how of manners of givenness”17 of phenomena in the intuitively given field of inquiry. Following Husserl, my focus here is to show how a transcendental phenomenological rational approach discloses horizons of indeterminacy arising at the limits of even a clear understanding of objects as a normative element of rational understanding which brings into view interconnections of the meanings of objects with other horizons of meaning in the world horizon. Given the viability of transcendental phenomenology as a normative theory of reason and its disclosure of horizons of indeterminacy as integral to rational understanding, the way opens to the further disclosure of the integral interdependence and complementarity of theoretical and practical reason. The initial methodological aim of transcendental phenomenological inquiry is to attain ever-increasing degrees of evidential clarity for objects, and, in this regard transcendental phenomenology remains, in a certain sense, teleological;18 yet, in the process of inquiry into evidence for objects in the world horizon, phenomenological inquiry leads to the disclosure of horizons of indeterminacy surrounding objects, which places a structural limit on the degree of clarity attainable for understanding. In Ideas 1, in a discussion of our understanding of physical things from the perspective of the natural attitude, Husserl characterizes the tension between the pursuit of a clear understanding of objects and the disclosure of horizons of indeterminacy surrounding objects in terms of “determinable indeterminateness,” suggesting that what remains indeterminate about an object could, in principle, be made determinate by turning the focus of attention towards the indeterminate and, thereby, making it determinate.19 However, in later works, such as Formal and Transcendental Logic and The Crisis of European Sciences and Transcendental Phenomenology, Husserl’s understanding of the significance of horizons of indeterminacy for phenomenological inquiry deepens to include reference to what he calls “non-predicative evidence”20 which points to the ultimate sources of meaning, including the meaning of the historically sedimented intersubjective constitution of subjectivity, which can be partially uncovered through a genetic phenomenological method of inquiry.21 In Formal and Transcendental Logic (originally published in 1929), Husserl explains the phenomenological process for evaluating evidence in terms of a judgment schema according to which the level of evidential fulfillment for an object in the forming of judgments is placed on a scale of ascending degrees of clarity. Building on and modifying the Cartesian idea of clear and distinct understanding as the criteria for true judgment, Husserl sets out possible degrees of intuitive clarity for understanding as vague, distinct, or clear.22 He describes vague judging as a judgment in which the evidence for the object is unclear and the object is not adequately understood but rather “meant only expectantly.”23 Next on the scale, distinct judging is described as judging with evidence, wherein the object of judgment, instead of merely

188  Robert H. Scott being meant, “now is properly and itself given.”24 Finally, a clear judgment pertains to “a givenness originaliter of the affairs themselves.”25 Once evidential fulfillment for the object of judgment reaches the stage of clarity, the judgment is considered by Husserl to have achieved adequate evidential fulfillment and may be taken as correct and true. It may seem that, at the stage of adequation and clarity, inquiry into the meaning of the object has come to an end; however, Husserl discovers that the meaning of an object does not end with clarity of evidence for it. A key insight of a transcendental phenomenological approach to rational understanding consists in its disclosure of the meaning of an object as extending beyond clear evidential determinations in view of horizons of indeterminacy that persist at the limits of clear understanding and are integral to the understanding of any object. In view of horizons of indeterminacy, even a judgment based on clear evidence remains open to questions such as: does the correctness or determinacy of the judgment exhaust the meaning of the object judged? Does the clarity of evidence guarantee that the determinative judgment is final, or that instrumentally or economically beneficial uses of the object, based on the judgment, need not be reviewed for possible harmful effects? Such ethico-practical questions arise in view of horizons of indeterminacy that surround the object and bind the meaning of the object to other horizons of significance in the world-horizon. One vital advantage of a transcendental phenomenological approach is that it reveals that such questions are integral to a thorough rational inquiry and, further, that ethico-practical concerns and demands are integral both to rational understanding and to the full meaning of objects. In response to the aforementioned questions, transcendental phenomenological reason will answer “no”—correctness does not exhaust the meaning of objects; even a clear judgment could be falsified if sufficient new evidence contradicts it, and, further, the implications of meaning for other persons and things in the world horizon are integral to the meaning of the object. In this manner, horizons of indeterminacy, disclosed through a normative rational approach, stand as limits to determinate understanding that raise demands for ethical responsibility in relation to objects through their wide-ranging interconnections of meaning with other horizons of significance in the world horizon. Not only do horizons of indeterminacy tie the meaning of objects to other horizons of significance in the world horizon, they also serve to bridge what Husserl refers to as the static method of phenomenology and the genetic method, developed in his later work (after 1919).26 Where static phenomenology focuses on applying the basic formal phenomenological techniques of the reduction and evaluation of evidence in the world horizon through intentional analysis, genetic phenomenology opens pathways towards understanding the temporal, historical, and intersubjective development of the meaning of self-identity, objects, and the world.27 Through genetic phenomenology, the persistent horizons of indeterminacy, which become evident through static phenomenological inquiry, now indicate not only the

Indeterminacy as Key to Reinterpretation  189 intertwining of the meaning of objects with other synchronous horizons of significance in the world horizon but also point to what Husserl calls the “non-predicative” historical and intersubjectively constituted diachronic sources of evidence for the world horizon. That is, even where clear evidence for an object has been attained, horizons of indeterminacy indicate, by way of genetic phenomenological inquiry, that predicated evidence can be traced back further to non-predicative historical and intersubjective transcendental sources. In this manner, horizons of indeterminacy surrounding objects point to what Husserl describes as non-predicative evidence for sedimented sources of significance. In Formal and Transcendental Logic, Husserl describes non-predicative evidence as evidence for the ultimate intuitive sources of predicative evidence. The ultimate sources of evidence, he explains, include the temporal and intersubjective constitution of subjectivity, also referred to as transcendental intersubjectivity.28 The highest level of transcendental phenomenological rational inquiry, therefore, leads to the disclosure of horizons of indeterminacy which, in addition to raising ethico-practical demands and indicating interconnections of meaning between objects of inquiry and other horizons of significance in the world horizon, point to non-predicative evidence for the transcendental and historical, intersubjective constitution of subjectivity. Referring to the non-predicative (or indeterminate) aspect of experience as the “first thing in the theory of evident judgments,” Husserl writes, the intrinsically first thing in the theory of evident judgments (and therefore in judgment-theory as a whole) is the genetical tracing of predicative evidences back to non-predicative evidence called experience.29 Shortly thereafter, he reiterates, the intentionality of predicative judgments leads back ultimately to the intentionality of experience.30 By my reading, the “intentionality of experience” here refers to an irreducible aspect of indeterminacy encountered at the limits of phenomenological clarity that points to both links between the meaning of the object with further horizons of significance in the world horizon and to the deep historical, transcendental, and intersubjective ultimate sources of meaning.31 Given that there is always a horizon of indeterminacy relevant to a thorough rational understanding of an object, the method of transcendental phenomenology discloses a demand for epistemic modesty in relation to the kind of clarity attainable by understanding, and, no less importantly, calls for the continuation of inquiry, in a largely ethico-practical mode, beyond determinate clarity to the wide-ranging, diachronic, and intersubjective implications of meaning as integral to knowing.

190  Robert H. Scott Another way into a phenomenological consideration of the basic meaning of horizons of indeterminacy is through a consideration of what Husserl describes as the internal and external horizons of an object. On the one hand, Husserl explains individual things as having what he calls an open “internal horizon” consisting of the evidence determinative of our understanding of the sense of the thing (which may become clear); on the other hand, each thing also has an open, external horizon of indeterminacy at the limits of the internal horizon which links the significance of the object to other things in the world. In Crisis, Husserl writes, just as the individual thing in perception has meaning only through an open horizon of “possible perceptions,” insofar as what is actually perceived “points” to a systematic multiplicity of all possible perceptual exhibitings belonging to it harmoniously, the thing has yet another horizon: besides this “internal horizon” it has an “external horizon” precisely as a thing within a field of things; and this points finally to the whole “world as perceptual world.”32 This passage indicates that while the internal horizon of sense for an object, while open, may become determined with clarity—as in the case of mathematical objects, for instance—the external horizon of the object maintains an aspect of indeterminacy that opens the meaning of the object to interconnections with other horizons of significance in the “field of things” that surround it.33 While Husserl does not dwell on the ethico-practical implications of the indeterminacy of the external horizon of objects, I contend that an important part of the meaning of objects that arises with external horizons of indeterminacy includes ethico-practical demands for responsibility both towards things themselves and in our dealings with things in relation to their both short- and long-term implications of meaning in relation to other things in the world-horizon. In view of the transcendental phenomenological theory of reason presented thus far, I now turn to retrieve the Aristotelian theoretical intellectual virtues of intuitive reason, scientific knowledge, and philosophic wisdom in order to consider how our conception of these intellectual virtues would change when reinterpreted through the lens of a transcendental phenomenological theory of reason.

A Transcendental Phenomenological Retrieval of Aristotle’s Theoretical Intellectual Virtues Intuitive Reason (noûs) In section 6.6 of Nicomachean Ethics, Aristotle defines the task of intuitive reason as that of grasping first principles.34 Leading up to this definition, he starts with two premises about first principles: 1) that they are invariable

Indeterminacy as Key to Reinterpretation  191 objects of knowledge, and 2) that they are not arrived at through deductive demonstration. Based on the first premise, he infers that first principles cannot be grasped through the practical virtues of art or practical wisdom, since both of these are concerned with variable objects. Based on the second premise, he infers that first principles cannot be grasped through science because the latter always involves deductive demonstration; for the same reason, he eliminates philosophical wisdom as the virtue exclusively dedicated to grasping first principles, since the latter includes scientific knowledge as integral to it. Through a kind of reductio, then, Aristotle assigns the task of grasping first principles to intuitive reason. In Nicomachean Ethics, however, Aristotle does not provide a detailed description of how intuitive reason goes about grasping first principles.35 In embarking on a creative retrieval of intuitive reason through a transcendental phenomenological approach, it is striking to note that the project of transcendental phenomenology as a whole may be characterized as a theory of intuition.36 Both Aristotle and Husserl associate intuitive reason with some version of what Aristotle calls, “the principle of the principles,”37 and this, for Aristotle, is due to the fundamental role that intuitive reason plays in grasping first principles. There is, however, an important difference between Aristotle and Husserl’s views on the role of intuition, and that is that Husserl expands the role of intuition by giving central place to the “intuitive sphere” as the general field of phenomenological inquiry. In Ideas 1, Husserl describes what he refers to as the “principle of all principles” for phenomenology as affirming, that every originary presentive intuition is a legitimizing source of cognition, that everything originarily (so to speak, in its “personal” actuality) offered to us in “intuition” is to be accepted simply as what it is presented as being, but also only within the limits in which it is presented there.38 In Husserl’s last book, Crisis, the centrality of the sphere of intuition continues to be upheld in his formulation of the transcendental reduction of experience to what he calls the lifeworld. Through the transcendental or “universal epoché,”39 the field of inquiry becomes the manifold transcendental phenomenon of the lifeworld as it is given to intuition, and the rational method of inquiry into the lifeworld involves intentional analysis, evaluation of evidence in terms of degrees of clarity, and a wide-ranging consideration of “the how of the world’s manners of givenness”40 that includes accounting for the temporal, historical, and intersubjective constitutive elements of evidence. In this manner, intuitive reason takes on a much broader and integral role for transcendental phenomenology than it has for Aristotle insofar as the domain of intuitive reason is now understood as synonymous with the manifold field of inquiry. Given the centrality for Husserl of what may be called intuitive reason, a transcendental phenomenological reinterpretation

192  Robert H. Scott of the Aristotelian intellectual virtue of intuitive reason will consider the latter to be an all-encompassing intellectual virtue insofar as intuition concerns the entire field of phenomenological inquiry. While Husserl does not go into depth in addressing the ethical implications of the practice of transcendental phenomenological inquiry into the intuitive sphere, he does indicate that, by virtue of the transcendental phenomenological recognition of the manifold of interconnections of meanings in the life-world,41 phenomenological investigation can have a transformative effect on the character of those who practice it. In regard to this effect, he writes, Perhaps it will even become manifest that the total phenomenological attitude and the epoché belonging to it are destined in essence to effect, at first, a complete personal transformation . . . (which) bears within itself the significance of the greatest existential transformation which is assigned as a task to mankind as such.42 While the practice of transcendental phenomenology begins, for Husserl, as a theoretical exercise, it is striking that Husserl here suggests that it can have a profound personal and practical transformative effect. Although he does not elaborate in detail on the sort of personal transformative effect transcendental phenomenology may have on its practitioners, an important implication of his claim here is that if the practice of intuitive reason through transcendental phenomenological inquiry contributes to the development of moral character then it must involve not only theoretical reason but also practical reason and, further, that the development of a kind of practical wisdom must be integral to the highest level of phenomenological reasoning. From a phenomenological perspective, then, intuitive reason must not be an exclusively theoretical virtue, as Aristotle understood it to be; rather, as re-interpreted through the lens of transcendental phenomenology, intuitive reason must involve an overlapping and complementary exercise of both theoretical and practical reason. Scientific Knowledge (epistéme) In Aristotle’s account, the virtue of scientific knowledge is concerned with the deduction of unchanging truths from first principles, meaning that the objects, truths, and meaning attainable through the activity of science are presupposed to be invariable and fully determinate. Without referring to Aristotle, Husserl provides clear guidance for how the intellectual virtue of scientific knowledge would be modified in view of a transcendental phenomenological theory of reason. The first problem with Aristotle’s account of scientific knowledge, highlighted by a transcendental phenomenological approach, arises in the setting up of a sharp separation between the exercise of intuitive reason and the deductive process of scientific knowledge. The

Indeterminacy as Key to Reinterpretation  193 problem with this sharp separation is that it allows for a reductive understanding of science as a narrowly deductive enterprise. By contrast, a transcendental phenomenological understanding of science discloses the activity of science as part of the universal problem of intuitive reason, not a narrowly deductive process.43 This transcendental phenomenological reinterpretation of the activity of science opens the way to a richer understanding of the practices and results of science as inseparable from intuitive sources of meaning in the lifeworld and wide-ranging interconnections of meaning with other horizons of significance in the world horizon. Understanding the practice of science as seamlessly connected with the universal problem of intuitive reason brings into view the demand for science to maintain or restore attentive awareness to horizons of indeterminacy at the limits of the clear, determinate understanding which bind the determinate insights of science with other horizons of meaning in the world horizon and which convey ethico-practical demands. In this manner, a transcendental phenomenological reinterpretation of the intellectual virtue of scientific knowledge serves to ensure that a full understanding of science, both in regard to its procedures and results, must include an account of the interconnections of the meaning of its objects with ethico-practical demands as well as other horizons of significance. Conversely, the seamless link between the intellectual virtues of intuitive reason and science serves as a reminder to non-scientists that there is a demand to take scientific findings seriously and that further ethical demands arise where the determinate findings of science have bearing on habits of action and practical decision making.44 I now turn briefly to consider the manner in which the separation of science from its intuitive sources, evident in Aristotle, increases in the modern age, in order to show that the relevance of the transcendental phenomenological critique of science as a “partial problem” within the broader problem of human meaning in the life-world remains just as relevant to the modern conception of science as it does to the Aristotelian conception. Despite modern refinements of the empirical scientific method, Husserl argues persuasively that the problematic separation between science and its intuitive sources in the life-world actually widens in the modern era. In Crisis he argues that, beginning with Galileo’s mathematization of nature, modern science loses track of its ultimate sources in the intuitive sphere of the life-world and becomes a “residual concept,”45 separated off from the manifold field of human meaning. While Galileo is rightly known as a great early modern mathematician and astronomer, what is less well known and what Husserl brings to light in the Crisis is that Galileo was also profoundly influential, in a negative way, in the development of a mode of scientific inquiry that neglects and fails to recognize the importance of the intuitive sources of scientific knowledge and the embeddedness of the practice of science in the interconnected, meaning-rich context of the lifeworld. While Galileo was, without question, brilliant in his development of original mathematical formulae and geometrical models to explain planetary motions

194  Robert H. Scott (among other contributions), he passed over the importance of the intuitive processes and sources from which mathematical idealities are drawn and in which they are embedded with other horizons of meaning. Consequently, from Galileo, Husserl argues, there arose, the surreptitious substitution of the mathematically substructed world of idealities for the only real world. . . . This substitution was promptly passed on to his successors, the physicists of all the succeeding centuries.46 Hence, despite his brilliance and deserved place of distinction in the history of science, Galileo failed to recognize the significance of the intuitive processes and sources from which mathematical idealities are formed and in which they are embedded with other horizons of meaning which have bearing on the full meaning of objects.47 Husserl’s critique of Galileo in Crisis suggests that the forgetfulness or neglect of the intuitive sources paved the way, in the modern age, for the conflation of nature with mathematical idealities, evident in the modern characterization of nature and being in terms of a “mechanical universe” which leads to a view of nature as deficient in human meaning and moral value. Since Galileo, as Husserl argues, there has been a tendency to conflate mathematical idealities with true being or the ultimate meaning of being, and where this occurs the method of mathematical science, its formulae, and the meaning of its results have tended to be understood in a limited, narrow way that focuses on the determinate aspects of scientific phenomena and neglects the wide-ranging interconnections of meaning in the lifeworld that constitute the existential and moral meaning of the world. Such neglect of the human meaning of science has, arguably, contributed to contemporary environmental problems insofar as they have resulted from historical and contemporary neglect of the long-range ethical and practical implications of human action in its interactions with the nonhuman environment. Another way of stating the problem is that, in the modern era, science lost its way insofar as it came to view itself narrowly as the practice of specialized skills (techné) in the service of technological development, losing sight of the embeddedness of its practices and results in the manifold and temporally extended field of meaning that includes both short- and long-term ethico-practical demands. In the foregoing transcendental phenomenological retrieval of the intellectual virtue of scientific knowledge, we do not disparage the value of all the discoveries of new algorithms, insights, and formulae and the development of advanced instruments accrued through modern science. However, a transcendental phenomenological understanding of science discloses such a narrowly focused construal of science as a partial problem within the larger universal problem of rational inquiry into the meaning of the lifeworld. As such, transcendental phenomenology recalls that the meaning of science does not end with mathematical idealities and empirical results;

Indeterminacy as Key to Reinterpretation  195 rather, the virtue of scientific activity, understood through the lens of transcendental phenomenology, calls for renewed and heightened awareness of the overlapping of the meanings of the practices and results of science with a wide range of other horizons of significance in the world horizon and of the ethico-practical demands that arise through this activity. As such, a transcendental phenomenological reinterpretation of the virtue of scientific knowledge serves as an emphatic reminder of the subjective correlates and intuitive sources of the mathematical idealities and empirical results of science along with the interconnections between the open external horizons of scientific insights with other horizons of significance in the lifeworld. I now turn, finally, to a transcendental phenomenological retrieval of the highest Aristotelian intellectual virtue: philosophic wisdom. Philosophic Wisdom (Sophía) In section 7 of book 6 of Nicomachean Ethics, Aristotle explains the intellectual virtue of philosophic wisdom as the highest form of theoretical reason that involves combined excellence in intuitive reason and scientific knowledge. In the foregoing transcendental phenomenological reinterpretation of intuitive reason and scientific knowledge, the practice of science has been reinterpreted and clarified as a basic, integral activity of transcendental phenomenological reason and may be understood as consisting of a focused yet wide-ranging phenomenological investigation of the various scientific areas of study within the manifold field of the world horizon. Understood in this way, a transcendental phenomenological approach to scientific work will include attentive awareness to the embeddedness of scientific inquiry in the manifold context of the lifeworld, and, regardless of how specialized the investigation may be, it will involve active interest in following the implications of meaning and ethical demands that arise through horizons of indeterminacy at the limits of clear understanding that emanate from both the procedures and results of scientific investigation. The primary sticking point between Aristotle’s conception of philosophic wisdom and a transcendental phenomenological conception of it centers on the notion of invariable objects and Aristotle’s privileging of complete determinacy as the primary goal of philosophic wisdom. Aristotle considers invariability to characterize the highest kind of object which he further describes as necessary, ungenerated, and imperishable.48 That is to say, Aristotle considers the objects of theoretical reason to be superior because they are invariable, and that is because they bear likeness to the pure immaterial substance of the Unmoved Mover. As noted earlier, by setting aside Aristotle’s onto-theological presuppositions, a transcendental phenomenological theory of reason dissolves the motivation for privileging invariability and complete determinacy as the highest goals of reason. To the contrary, a transcendental phenomenological approach discloses the embeddedness of the meaning of scientific objects in the intuitive context of the lifeworld

196  Robert H. Scott and consequently that, at least as important as the determinate meaning of objects is their meaning in relation to other horizons of significance in the world horizon, especially ethico-practical horizons of meaning, which arise from horizons of indeterminacy surrounding objects. In view of these transcendental phenomenological insights, we may reinterpret philosophic wisdom as a kind of thinking about thinking that involves, rather than contemplation of invariable objects, attentive awareness to horizons of indeterminacy surrounding objects and of the interconnections of the procedures and results of science with a multiplicity of other horizons of meaning in the world horizon. The expanded awareness of horizons of indeterminacy and of the wide-ranging significance of concepts and things that a transcendental phenomenological reinterpretation of philosophic wisdom reveals, points to the necessary involvement of practical wisdom or a kind of deliberative, dialogical, and discursive rationality, along with the theoretical activity of determining sense, as integral to the highest level of rational activity. In this manner, a transcendental phenomenological re-conception of philosophic wisdom discloses the necessary overlap and interdependence of practical and theoretical reasoning and deflates the hierarchical ranking, found in Aristotle, of theoretical reason over practical reason. It may even be said that philosophic wisdom, when viewed through the lens of transcendental phenomenology, must include practical wisdom insofar as the philosophic wisdom necessarily involves and demands, in tandem with excellence in theoretical, scientific determination of sense, excellence in the responsible and deliberative activity characteristic of practical wisdom.

Conclusion Seeking to identify what the human function is, in book 1 section 7 of Nicomachean Ethics, Aristotle presents a kind of reductio in which he sets out a series of human activities and then considers which of them differentiates humans from other entities. After eliminating living and perceiving as specific to humans, Aristotle concludes that the specifically human function must consist of an activity of the soul that follows a rational principle.49 Granting that the human function is, according to Aristotle, to act according to a rational principle, and given that things attain their highest end by performing their function well, it follows that human beings who act according to a rational principle to the highest degree, will put themselves in the best possible position to attain happiness. On this point, a transcendental phenomenological account may agree with Aristotle, and it will further agree on the idea that identifying and cultivating the highest level of rational activity is a central concern for ethical theory and practice. However, the application of a transcendental phenomenological theory of reason will depart from Aristotle on his view that the highest level of rational activity consists of a strictly theoretical form of reason aimed at attaining fully determinate knowledge of invariable objects.

Indeterminacy as Key to Reinterpretation  197 The central insight I hope to have conveyed through this transcendental phenomenological reinterpretation of the intellectual virtues is that philosophic wisdom is not limited to isolating determinate attributes of objects; rather, the best use of reason must entail, along with the theoretical determination and clarification of concepts and objects, recognition of horizons of indeterminacy that surround objects and the ethico-practical demands that arise from them along with other wide-ranging interconnections of meaning in the world horizon. Finally, drawing together Aristotle’s ideal of virtue as excellence in fulfilling the human function and Husserl’s suggestion that the exercise of transcendental phenomenology can lead to personal transformation, I suggest that the transformation that may occur through excellence in rational activity, understood through the lens of a transcendental phenomenological theory of reason, has the potential to motivate global social and ecological renewal insofar as rational activity so understood entails recognizing and responding appropriately to both short- and long-term ethicopractical demands that arise from horizons of indeterminacy that emanate from both human and nonhuman entities in the world horizon.

Notes 1 Aristotle (2009, p. 12, 1098a 10–15). 2 The other two Aristotelian intellectual virtues, art (téchne) and practical wisdom (phrónesis) are virtues of practical or deliberative reason and will not be addressed in detail here. 3 Husserl defines the world horizon as the “horizon of possible thing-experience” (1970, p. 138), and he goes on to state that objects are always given “in principle only in such a way that we are conscious of them as things or objects within the world-horizon” (1970, p. 143). Thus, the world horizon is the necessary, broad horizon within which phenomena appear. As Gail Weiss notes, the world horizon (and the overlapping, significant interconnections that arise within it) “tends to be rendered invisible” when one focuses exclusively on the object of attention. One aim of transcendental phenomenology, she further notes, is to counter this tendency. Weiss describes the world horizon as the general horizon that “situates the perceptual field” (2008, pp. 12 and 29, note 4). A great asset of phenomenology is that it insists on preserving awareness of the world horizon as the context within which objects appear, as well as of the interconnections of meaning that arise through the intertwinement of horizons of meaning in the world horizon. 4 Aristotle (2009, p. 104, 1139b 17–18). 5 Ibid., p. 105, 1140a 1–1140b 12. 6 Ibid, pp. 104–105, 1139b 18–35 and pp. 107–108, 1140b 31–1141b 3. 7 Ibid., p. 196, 1178a 5–11. 8 See Aristotle (2002, 1072 b 7–31) on the Unmoved Mover. For Aristotle’s discussion in Nicomachean Ethics of the happiest human life as one characterized by activity that most closely approximates to the activity of the Unmoved Mover, see Aristotle (2009, p. 197, 1178b 21–24). 9 For example, architectural objects aim at being both beautiful and useful, and one could achieve this through various forms of architecture (neo-classical, art deco, craftsman, etc.). Similarly, for practical wisdom, how to apply the mean well can vary from one person to another. To illustrate this, Aristotle gives the example of eating, noting that, for Milo the Wrestler, the mean of how much to

198  Robert H. Scott eat will be much greater than for the average person. See Aristotle (2009, p. 30, 1106a 29–1106b7). 10 Ibid, pp. 102–103, 1139a 6–12. 11 Ibid., p. 197, 1178b 8–23 and p. 196, 1178a 9–10. 12 In Ideas 1, Husserl indicates that, under the phenomenological reduction, the idea of God as a transcendent being is bracketed out from the field of inquiry. Referring to God as the “ ‘absolute’ and ‘transcendent’ being,” he writes, “we extend the phenomenological reduction to include this ‘absolute’ and ‘transcendent’ being.” Husserl does, however, remain quite open to the possibility of the development of phenomenological theology through inquiry into “modes in which transcendencies are made known.” See Husserl (1982, pp. 134 and 117). 13 Husserl (1970, p. 143). Also, Weiss (2008, p. 29) discusses the rich significance of indeterminacy in relation to internal and external horizons in Husserl. She notes that the significance of indeterminacy has been developed in terms of ambiguity (by DeBeauvoir), absurdity (by Camus), and reversibility (by Merleau-Ponty), and she goes on to explore the manner in which horizons of indeterminacy play an important role in the formation and reformation of attitudes towards the body, gender, race, and class. In this chapter, I suggest that the significance of indeterminacy extends to include ethico-practical demands to both human and nonhuman entities. 14 This text will be referred to, in what follows, as Crisis. 15 Husserl (1970, pp. 12–13). 16 As Dermot Moran notes, Husserl refers to several different types of reduction and, citing Iso Kern, that there are several different ways to enter the reduction. See Moran (2000, pp. 146–147) and Kern (1977, pp. 126–149). The various reductions, in my view, can be understood as steps on the way to the universal, transcendental phenomenological reduction to the intuitive sphere of inquiry which may be described most broadly as the world horizon. In support of my view, Husserl refers to a “multiplicity of steps, each of which has, in a new way, the character of an epoché” which lead to the transcendental epoché. See Husserl (1970, p. 135). 17 Ibid., p. 263. 18 Hence, transcendental phenomenological reason may be considered teleological in more of a Cartesian than Aristotelian sense in that the phenomenological gathering of evidence for objects aims at gaining clear and distinct understanding. However, for Husserl, the Cartesian notion of clear and distinct evidence must be qualified in that, for phenomenology (unlike for Descartes), clarity of understanding must include recognition of horizons of indeterminacy surrounding objects. 19 In Ideas 1, in a discussion of inquiry from the perspective of the natural attitude, Husserl writes: “Necessarily there always remains a horizon of determinable indeterminateness, no matter how far we go in our experience, no matter how extensive the continua of actual perceptions of the same thing may be through which we have passed.” See Husserl (1982, p. 95). 20 Husserl (1969, p. 249). 21 For more on Husserl’s development of the genetic method of phenomenology, see Donohoe (2016). 22 Husserl (1969, pp. 56–62). 23 Ibid., p. 56. 24 Ibid. 25 Ibid., p. 61. 26 For more on the transition from static to genetic phenomenology in Husserl’s work, see Donohoe (2016, especially pp. 30–37).

Indeterminacy as Key to Reinterpretation  199 7 Ibid., p. 21 2 28 Husserl (1969, p. 249). 29 Ibid., pp. 209 and 223. 30 Ibid., p. 210. 31 While it extends beyond the bounds of this chapter to develop the importance of transcendental intersubjectivity, it is important to note that the interconnection between the ego and a wide-range of others in the world horizon not only indicates the inherently social character of the ego, it also points to the historical, diachronic intersubjective constitution of the ego, evident in its acquisition of language and culture. For more on the relation between genetic phenomenology and transcendental intersubjectivity, see Donohoe (2016, pp. 71–110). 32 Husserl (1970, p. 162) 33 As noted earlier, Weiss points out that there is an aspect of indeterminacy in both the internal and the external horizons of an object. See Weiss (2008, p. 29). In my view, the indeterminacy of the external horizon of things conveys ethicopractical significance though its role in interconnecting the meaning of objects to other horizons of significance which include ethico-practical horizons of significance. 34 Aristotle (2009, pp. 107–108, 1140b 30–1141a 8). 35 In the last section of Posterior Analytics (II.XIX), Aristotle describes in more detail the process by which noûs grasps first principles. (I am grateful to Greg Moss for directing me to this passage in Aristotle.) In that section, Aristotle indicates that noûs employs sense perception and induction to arrive at first principles, and he is (as in Nicomachean Ethics) careful to distinguish the work of noûs from scientific demonstration which operates through deduction from first principles. See Aristotle (1901, pp. 126–128). 36 See, for example, Emmanuel Levinas’ discussion of phenomenology as a theory of intuition in Levinas (1995, pp. 65–151). 37 Aristotle (1901, p. 128). 38 Husserl (1982, p. 44). 39 Husserl (1970, p. 156). 40 Ibid., p. 160. 41 Husserl’s recognition of the importance of interconnections of meaning in the intuitive sphere of inquiry is evident, for example, in the title of section 48 of Crisis: “Anything that is—whatever its meaning and to whatever region it belongs—is an index of a subjective system of correlations.” Ibid. p. 165. 42 Ibid., p. 137. 43 Ibid., pp. 9 and 134–135. On these pages, Husserl makes a similar critique in relation to modern science which, he argues, by detaching itself from its intuitive sources has reduced itself, unnecessarily, to a “residual concept” (9) and a “partial problem” (135) and, as such, tends to neglect important aspects of meaning. 44 For example, the findings of climate science on anthropogenic climate change carry important ethical demands both to recognize the validity and seriousness of the findings and respond with appropriate changes in attitudes, judgments, and habits. For more on the moral responsibility that arises with the findings of climate science, see Scott (2018). 45 Husserl (1970, p. 9). 46 Ibid., pp. 48–49. 47 Ibid., pp. 49 and 162. 48 Aristotle (2009, p. 104, 1139b 22–25). 49 Ibid., pp. 11–12, 1097b 28–1098a 18.

200  Robert H. Scott

Works Cited Aristotle. 2002. Metaphysics, translated by Joe Sachs. Santa Fe, NM: Green Lion Press. ———. 2009. Nicomachean Ethics, translated by David Ross. Oxford: Oxford World Classics. ———. 1901. Posterior Analytics, translated by E. S. Bouchier, B.A. Oxford: Blackwell. Donohoe, J. 2016. Husserl on Ethics and Intersubjectivity: From Static to Genetic Phenomenology, 2nd ed. Toronto: University of Toronto Press. Husserl, E. 1969. Formal and Transcendental Logic, translated by Dorion Cairnes. The Hague: Martinus Nijhoff Publishers. ———. 1982. Ideas Pertaining to a Pure Phenomenology and to a Phenomenological Philosophy, First Book, translated by F. Kersten. Dordtrecht: Kluwer. ———. 1970. The Crisis of European Sciences and Transcendental Phenomenology, translated by David Carr. Evanston, IL: Northwestern University Press. Kern, I. 1977. “The Three Ways to the Transcendental Reduction in the Philosophy of Edmund Husserl,” in P. McCormick and F. Elliston (Eds.), Husserl: Expositions and Appraisals, Sound Bend, IN: Notre Dame University Press. Levinas, E. 1995. The Theory of Intuition in Husserl’s Phenomenology, translated by André Orianne. Evanston, IL: Northwestern University Press. Moran, D. 2000. Introduction to Phenomenology, New York, NY: Routledge,Taylor & Francis, Inc. Scott, R. H. 2018. “A Phenomenological Theory of Ecological Responsibility and Its Implications for Moral Agency in Climate Change.” Journal of Agricultural and Environmental Ethics (forthcoming). Available online at https://doi.org/10.1007/ s10806-018-9713-z. Weiss, G. 2008. Refiguring the Ordinary. Bloomington: Indiana University Press.

9 The Effability of the Normative Todd May

There are many types of indeterminacy. There is perceptual indeterminacy: is that a ship I see in the distance or a lighthouse on a promontory? There are indeterminacies of number: is my weight 148 or 149 pounds, since the scale keeps going back and forth? There are indeterminacies of taste: do I like pizza better than I like ice cream, or vice versa? There are also deeper indeterminacies, ones that resist all forms of resolution. After all, I might be able to resolve these different indeterminacies by moving closer to the object or getting a more accurate scale or thinking about which I would prefer to have if I could only have one for the rest of my life. These deeper indeterminacies, we might say, are structural. It is in the nature of the phenomenon to be indeterminate. Prominent among these structural indeterminacies is the phenomenon of ineffability. Ineffability is where some region of experience cannot be rendered conceptually. It can be a particularly vexed form of indeterminacy, since it at once seems to require and to refute attempts to understand it. Ineffability is in the air these days, and has been for some time. In many areas of Continental philosophy, it is the very ethos in which thought is conducted. Think, for instance, of Emmanuel Levinas’s claim that we are haunted by the irreducible infinity of the other, Jacques Derrida’s absence that cannot be reduced to presence, Gilles Deleuze’s positing of a virtual field of difference, and more recently Adriana Cavarero’s claim that philosophy has dissolved the singularity of human beings expressed in their voices to general conceptual categories. It is not for nothing that Gary Gutting categorizes recent French philosophy as a project of “thinking the impossible.”1 Hand in hand with an orientation toward the ineffable is a critique of what might be called linguistic reductionism, which is roughly the idea that we can capture experience or being or others adequately by means of linguistic categories. This view involves a particular conception of language and a particular conception of categories, or better a conception of language that sees it primarily in terms of categories. We should be careful not to view the critique of linguistic reductionism as a critique of language tout court, mostly because there are those among the thinkers canvassed a moment ago, Derrida in particular, who refuse to see language in terms of

202  Todd May strictly delineated categories. What is sought in what is often referred to as “the limits of language” is not so much the limits of language itself, but instead the limits of language conceived in terms of more or less strictly delimited conceptual categories. (For the purposes of this chapter, I will not discuss the relationship of the linguistic to the conceptual, but will simply assume that they are closely linked.) This limitation, with reference to recent thinkers such as Nancy, Malabou, Stiegler, and Marion, has been recently been summed up by Ian James as a rejection of what he calls the “linguistic paradigm” “in the name of a systematic attempt to radically rethink questions of worldliness, shared embodied existence and sensible-intelligible experience.”2 James’s view is that what has commonly come to be called “the linguistic turn” in philosophy—a turn that affected both Continental and Anglo-American philosophy, has been coming under criticism in recent French thought for its reduction of both ontology and experience to the linguistic, to categories. Among the paradigmatic frameworks of recent thought, then, is an embrace of the ineffable that reveals the limits of linguistic reductionism in favor of such themes as singularity, vitalism, materiality, and otherness. What these alternatives have in common is their resistance to being rendered in conceptual terms, terms that would see them as adequately rendered through the proper linguistic categories. Singularity, for instance, cannot be rendered categorially because any category would allow the possibility of something else falling within its purview, which then would no longer be singular. Or again otherness, particularly in Levinas’s hands, cannot be brought into one’s own categories without losing the very fact of its being other. What I argue in this paper cuts directly against the orientation of this framework. I am going to claim that the realm of the normative is deeply linguistic, and moreover that this is a good thing. In contrast to the attempt of many thinkers to remove the normative from the conceptual or the linguistic, I would like to show that it is central to normativity to have a linguistic reference, a reference rooted precisely in the sense of conceptual categories that so concern thinkers of the ineffable. My claim is not that normativity can be reduced to conceptual categories. In fact, I think that some simple examples will show that it cannot. In that sense, I don’t argue that thinkers of the ineffable are mistaken. Rather, it is that an important— indeed crucial—aspect of normativity is neglected when we focus on the ineffable. At the end of the chapter, I will use Adriana Caverero’s thought of the politics of the voice to illustrate this. This is not because I find her views a particular egregious example of such neglect. Instead, it is because I find her views among the most promising of recent discussions appealing to the ineffable, and want to show the importance of recognizing the role that linguistic articulation plays in buttressing those views. Indeed, it is in answer to a question she raised to me several years ago at a conference at the New School that this issue came alive for me. This chapter can be seen as an extended version of my response to her question.

The Effability of the Normative  203 At the outset, it would be worth asking the simply stated question, what is normativity? If I am going to argue that normativity is deeply conceptual, then perhaps I ought to say something about the concept of the normative. At a first go, we can say that normativity is the realm of norms. What, then, are norms? That is a more difficult matter. Think of the variety of norms that exist in a variety of contexts. There are epistemic norms regarding seeking truth and offering justification; there are athletic norms involving training and good sportsmanship; there are moral norms regarding how people ought to be treated; there are norms of driving, artistic beauty, politeness, friendship, and even classroom behavior. All of these norms are, of course, contestable. One does not have to recognize or abide by them uncritically. However, if they are contested they are so in the name of other norms that are held to be more relevant, fruitful, exciting, interesting, valuable, or something else that is worth embracing. Norms appear not only in the guise of prohibitions and obligations, but also permissions, recommendations, enticements, discouragements, pleadings, warnings, many types of orienting, and even lots of nudging. They all have something to do with influencing behavior, and we might define them broadly as various standards for such influence. The term standards is doing a lot of work here, encompassing everything from strict prohibitions to mild encouragements. But the underlying idea it captures is that norms involve criteria of better or worse ways of going about things, whether those things be our behavior, our creations, our practices of knowing, our emotions, our relationships with others, or indeed our lives. It might seem that right off I have biased my reflections on normativity toward the conceptual. After all, if norms are standards, what can standards be except linguistically articulated rules? It would be a mistake, however, to reduce standards to rules. This is for at least two reasons. One is the famous Wittgensteinian argument regarding the potential infinite regress of rulefollowing.3 If all norms were rules, then we would need standards of success for following those rules, which would itself require rules, which would in turn require standards of success, etc. Correct rule-following, then, cannot be a matter of having categorial or conceptual standards all the way down. At some point, rules must give way to ways of doing things, ways of “going on” that are recognized by those involved in the practice in which the rule appears. The second, more pedestrian reason that standards cannot be reduced to linguistic rules is that there are lots of norms that resist being stated in linguistic fashion. Think of learning to ride a bike. There are many norms that are difficult if not impossible to state linguistically, such as, “When the bike starts to bend that way, do this,” where this is some sort of corporeal motion. Or again, consider norms of politeness. There are ways to modulate one’s voice to display different levels and types of appreciation. How could we put them all in the form of rules? We all know people who seem abrupt or insincere when expressing appreciation for a favor we’ve done

204  Todd May for them, but it’s often difficult to say where exactly they went wrong, aside perhaps from some general remarks about saying thanks too quickly or not modulating the gratitude to the size of the favor. But there is no chart to tell us how to express appreciation for different kinds of favors. (And if there were, it would likely be self-defeating. How would we know that our favor really was appreciated if the appreciation was in accord with formal rules. It’s precisely the ineffability of the forms of appreciation that allow them to render gratitude sincerely.) The discussion here, I should note in passing, intersects with the analytic discussion of whether one can reduce all knowing-how to knowing-that, and my claim is that one cannot. Various sorts of skills that are normative for engaging in a practice or at least engaging in it well seem to me to resist reduction to linguistic form. There are those, such as Jason Stanley, who have argued that one can.4 If he turns out to be right, then there is even less ineffability to norms than I’m claiming, so I will leave this debate to the side. If so many norms concerning the skillful engagement in a practice are irreducible to linguistic or conceptual articulation, then is it possible that beneath the linguistic or the conceptual lay an ineffable normativity that perhaps grounds those articulations or at least resists them? Have Levinas, Derrida, Cavarero, and/or others cottoned on to an ineffable normative underpinning of our experience, one that linguistically articulated norms can at best only be founded upon and at worst simply betray? At times that seems to be the common framework of these thinkers’ approaches to normativity, although with the exception of Levinas it is never laid out in precisely those terms.5 In Derrida’s case, it is more nearly implied by his deconstructive approach, through which he seeks to show that categories that are said to be founding for a particular philosophical view require a category they seek to excise in order to function; thus there is an economic relationship between the two that cannot result in a stable concept. This is why, for instance, Derrida claims that “differance is neither a word nor a concept.”6 As far as Cavarero goes, we will discuss her own view presently. Inasmuch as these thinkers, and others like them, want to hold that the normative is ultimately ineffable, however, they are mistaken. To see this, let’s first return to the pedestrian examples of ineffable normativity. When learning to ride a bike, one might ask, “Why should I do this when the bike bends that way?” An answer might be something like, “If you don’t do that, there will be too much weight on one side and the bike will tip over,” or, in short form, “To keep from falling off.” One of the normative elements of skillful bike-riding involves following the norm that one actually stay on the bike. That norm is linguistically articulated. Now turn to the second example of politeness. “Why,” one might ask, “should I modulate my voice this way in order to express politeness?” And here the answer might be something like, “It shows real gratitude.” That would be the same idea as the bike-riding example.

The Effability of the Normative  205 But let’s take the politeness example another step. Suppose the recipient of this advice responds, “Maybe it does, but it also seems a bit patronizing to act like that.” What is happening here? A challenge, a normative challenge, is being issued to a norm. And note that this challenge, while linguistically articulated, is being issued to a non-linguistically articulated norm. While the challenged norm is a way of acting that resists articulation into any form of linguistic rule or standard, the challenge itself is linguistically articulated, and articulated in a way that involves a recognizable normative criticism, that of being patronizing. Is this possibility coherent? Of course it is. In the practice of polite interaction, there are various norms at play. Some of them are linguistically articulable, some not. How to modulate one’s voice to express gratitude in a particular situation resists linguistic articulation, but saying that one should, except in unusual circumstances, express gratitude for a favor does not resist such articulation—in fact, it was just articulated on the left side of the hyphen. In the case at hand, if we assume that it is not polite to patronize someone, then what we have is a challenge from within the practice of politeness (a culturally specific practice, to be sure) to certain norms of the practice. There is nothing mysterious about this. But if one can offer a coherent and relevant linguistically articulated challenge to a non-linguistically articulable norm, then we cannot assume that there is some layer of ineffable normativity outside or beyond or founding for linguistically articulated normativity. The reason for this is straightforward. If beneath all linguistically articulated normativity lies a layer of ineffable normativity, then no linguistically articulated challenge would have bearing on that ineffable layer. That layer would be grounding for any linguistically articulated normativity, and therefore immune to challenge from the layers above. One might be tempted to raise an objection at this point. Perhaps the practice of politeness is just a skewed example. Politeness might be said to be a superficial cultural practice that does not tell us anything deep about normativity. Perhaps the real normativity, the deep stuff, is ineffable, in contrast to politeness. It is not clear to me that politeness, although certainly culturally specific, is not an important aspect of social space. That it provides much of the glue that holds a society together is evidenced by its absence in our current political context. However, I will not insist on that point. Rather, I want to point out that what goes for practices of politeness also goes for other practices, as well. Consider here arguments about the role of emotion in moral theory and action. For those who hold that emotion is an important aspect of our moral relationships, there will likely be an ineffable aspect to moral practice, in contrast to those who want to see morality as more of a rational affair. However—and this is the point—the debate about emotion and reason in morality is one that is carried out through linguistic discourse. How else might it be had?

206  Todd May In fact, one can issue challenges to certain ineffable norms of a practice, or even to the normativity of an entire practice, on the basis of norms coming from other practices. Normative challenge or reflection does not need to remain intra-practice. One way to read certain works by Michel Foucault— and I would argue it is the right way—is to see them as involving (often implied) normative critique of particular practices on the basis of norms rooted in other practices. Discipline and Punish, for instance, involves a critique of norms of psychological normality, some of which resist linguistic articulation, on the basis of norms stemming from political discourse and even the practice of history. To be sure, Foucault himself was not always clear on this matter. However, his periodic invocations of the term intolerable marks its appearance in his thought.7 How, then, might we think of the relationship between the non-linguistically articulable norms and linguistically articulated ones? We should resist the simple model of a layer of non-linguistic norms that either rest beneath or upon a layer of linguistic ones. The above examples show the untenability of both options. Non-linguistic norms cannot serve as a ground or point of reduction for linguistic norms because linguistically informed reflection can criticize and alter them. However, to think of non-linguistic norms as derivable from linguistic ones involves collapsing knowings-how into knowingsthat. It is hard to see how to derive, for example, the proper way to ride a bicycle or show gratitude for a specific favor from a set of linguistically articulated norms. Moreover, the attempt at a derivation of norms from rules runs into Wittgenstein’s rule-following argument. If there are linguistically articulated normative standards all the way down, we face the threat of an infinite regress. There is yet another reason to resist the two-distinct-layer view: at what might be called the base level, there are both non-linguistic norms of the kind we have seen here and linguistically articulated norms. Let us take the admittedly anodyne and yet commonly accepted norm, “All things equal, do not kill innocent people.” Now there can be arguments about what constitutes innocence, and Wittgensteinian worries about what kinds of behavior meet this norm, but a norm like this has had a good bit of stability in orienting how we think it proper to go about our lives. To one extent or another, our behavior ought to answer to this norm. To see this is to recognize that the base level is neither simply a set of linguistically articulated norms nor an ineffable layer of responsiveness, but is a complex layer of both. The next level is one of linguistically articulable norms. This is because the next level consists in reflections upon, inferences from, questions about, and criticisms of the base level. This is where aspects or elements of the base level are held up for scrutiny. It is the level where one asks and perhaps answers questions like, “Why should I turn the bike this way when it bends that way?” and “Why should I modulate my voice this way in order to express politeness?” It is also the level in which one asks questions like what it is to be innocent, whether the innocent are never liable to be killed, and

The Effability of the Normative  207 what the difference is between killing someone and letting them die. This is the level at which Foucault’s work addresses us, although, as I mentioned, he himself rarely addresses us in such terms. It is crucial to recognize here that the second level, the level of normative reflection, although in one sense grounded in the first level, is not grounded in it in a way that would make it solely derivative from the first. The sense in which it is grounded in the base level is that without the base level there would be nothing for the second level to reflect upon. The second level involves linguistically articulable reflection upon certain norms, and therefore requires the existence of some norms to reflect upon. However, that reflection need not ratify or endorse those norms. Reflection might endorse the norms upon which it reflects or it might reject them. The base level, then, does not have any normative priority over the second level. If anything, there is a certain normative priority possessed by the second level, since it judges norms in the first level. However, one should not make too much of this. In making judgments about certain norms, reflection on them often (perhaps always) appeals to other norms at the base level and, therefore, does not stand normatively above the base level. Rather, there is an interaction in which the second level both judges and appeals to norms at the base level. Both elements—­ judgment and appeal—are necessary at the second level. Judgment is necessary because that is precisely what reflection is doing. It is judging— endorsing, challenging, questioning, nuancing—a norm or set of norms at the base level. Appeal is necessary as well because without other norms to appeal to the judgment could not happen as normative judgment. So rather than a strict normative priority residing in either of the levels, there is a normative interaction between the base and second levels. One might ask why there is not a third level that would involve considerations of aspects of the second level, and then not a fourth level, in a potential infinite regress. The answer is that there can certainly be a third level, and perhaps a fourth one as well, and so on. To claim that the levels necessarily come to an end somewhere would be to claim that foundationalism is true. This is because there would be a necessary stopping point, a point beyond which questions cannot be raised. To say that any particular level of reflection is the final one would be to claim that it offers the ultimate answers to our reflective questions and that those questions must come to a rest there. And that is one form—although not the only form—of foundationalism. Rather, with regard to levels, the situation is summed up by Wittgenstein’s phrase that, at some level of reflection, “If I have exhausted my justifications, I have reached bedrock and my spade is turned.”8 It’s not that there is an argument that one cannot ascend another level; it is that one just doesn’t know what to say any more. One’s reasons run out, not because the final answer has been given but because one no longer knows what to say. As Wittgenstein says elsewhere, “At the core of all well-founded belief, lies belief that is unfounded.”9

208  Todd May The last thing to note here is that both (or more) levels operate within the context of specific practices. This is a point I have addressed at length elsewhere and can only gloss here.10 Behavioral appropriatenesses, inferential relationships, normative judgments, and epistemic orientations are all developed at the level of specific social practices such as baseball, psychotherapy, philosophy, chess, computer programming, child-rearing, church-going, and house-building. The character of these practices orients and constrains both layers. They provide frameworks within which these various norms—both non-linguistic and linguistic, and at both levels—operate. These frameworks are not the final arbiters of norms. Not only can the norms in them be criticized, but the entire normative structure of a practice can also be criticized from the perspective of another practice, as we have seen that Foucault does. However, we should not think of norms either as free-floating, distinct from one another, or alternatively as all part of a seamless whole that might be called “human behavioral norms.” Instead, they arise within and refer to specific, often culturally inflected, practices. And this, I should mention in passing, is one of the reasons it is important to study other cultures or other practices distant in time from our own: to loosen the grip of naturalness that the norms of our practices often have upon us. With this picture of normativity in hand, I want to turn briefly to Adriana Cavarero’s book For More Than One Voice: Toward a Philosophy of Vocal Expression. The treatment of this book that I’m suggesting here might be equally applicable to Levinas’s work or Derrida’s later work on ethics and politics. My choice of Cavarero over these others is motivated by two concerns. First, especially with Levinas and at times Derrida, it is unclear whether their work is to be understood as normatively or descriptively motivated. Is Levinas, for instance, offering us a description of the way we, in fact, encounter others or commending to us a way of thinking about (or going about) approaching others? At times, his approach seems more structural than normative, although it seems motivated by a set of normative concerns. Cavarero’s position is more straightforwardly normative and so offers us a cleaner case to work with. Second, I think Cavarero’s work provides suggestive new areas for thought, and not only in the work I will discuss here. Her treatments of narrativity, of women in Plato’s work, and of terrorism and what she calls “horrorism” are rich discussions that deserve more attention than they have received up to now. Inasmuch as this piece can contribute to calling attention to her work, I would like to do so, even if my remarks here will involve criticism of one of her positions. In discussing For More Than One Voice, I am interested in focusing on its central claim. One way Cavarero puts this claim is this: “the ontological horizon that is disclosed by the voice—or what we want to call a vocal ontology of uniqueness—stands in contrast to the various ontologies of fictional entities that the philosophical tradition, over the course of its historical development, designates with names like ‘man,’ ‘subject,’ ‘individual.’ ”11

The Effability of the Normative  209 For Cavarero, the philosophical tradition—which she sometimes calls, in an echo of Heidegger and Derrida’s thought, metaphysics—neglects the uniqueness of human beings by subsuming them under general categories. Among the ways this neglect is evidenced is through the gradual effacement of the uniqueness of human voices in favor of the semantics of what is being said. She invokes Levinas’s distinction between the Saying and the Said as one way to approach this, but insists that the Saying is not simply the act of speaking; it is instead the individual voice of the person who speaks. As with her earlier book, Relating Narratives, Cavarero seeks to call our attention to the unique individual human beneath what she sees as a philosophical discourse that dissolves it in more-general categories. In her view, focusing on the individuality of the voice is one of the most promising ways to do so. One might see a similarity between her project and that of Levinas’s focus on the face, a similarity that she acknowledges. However, she is a bit chary about Levinas’s approach, since the face requires a more visual approach, and, as she puts the point, “this looking runs the risk of continuing to pay homage to the philosophical centrality of theoria. This ontology entrusted to the eye has been the lifeblood of western metaphysics since Plato.”12 By switching the locus of individuality or singularity from the gaze to the voice, Cavarero hopes to replace the superiority and distance associated with gazing upon another with the intimate presence of the voice. What would it be to respond to another’s voice? Cavarero utilizes the relation of mother and infant to illustrate such a response. Even before a child can speak, there is a sonorous relation—she sometimes describes it as a kind of singing—between the infant and its mother, one in which each recognizes the uniqueness of the other through the voice and responds to it with a voice of one’s own. This response cannot be a semantic one for the obvious reason that the infant cannot yet understand speech. Moreover, Cavarero argues, even when we do learn to immerse ourselves in the semantics of our language, our voices remained coiled within every sentence we speak. It is precisely her project to have us recognize this and learn to respond to that voice, because in doing so we respond to the individuality of the person speaking. In placing this project within the normative framework I have suggested, we should first note that the kind of response appropriate to a voice is not simply a linguistically articulable one. To be sure, to speak to another involves a semantics. Cavarero insists that her proposal is not to abandon semantics for the pure voice. “It is not a matter of feminizing politics; nor is it a question of making politics coincide with the pure voice by insisting on the subversive power of vocal pleasure. Rather, it is a matter of tracing speech back to its vocalic roots, extricating speech at the same time from the perverse binary economy that splits the vocalic from the semantic and divides them into the two genders of the human species.”13 However, the politics of such a recognition, even when it involves speaking a semantically structured language, happens in good part at a level that is not—and cannot

210  Todd May be—captured through a set of linguistically articulable norms. It is, instead, a matter of learning how to listen to the voice and respond to it with one’s own, a corporeal exercise that is more like learning to ride a bike than learning to understand a philosophical view. But what of Cavarero’s discourse itself, her discourse about the voice? She does not seek to call our attention to the voice through a set of ineffable gestures. Instead, she offers an articulate, a semantically grounded, analysis that shows the gradual effacement of the voice throughout the history of philosophy, calls our attention to the relation of voice to individuality, and urges us to recover a relationship to that voice. We might say that she argues for a particular norm: “Respond to the voice.” This norm is just as articulate, just as semantically grounded, as the norm “All things equal, do not kill innocent people.” What she offers is a second level critique of certain practices and norms that turn us away from the voice as well as an argument, on the basis of the recognition of the individuality of people, for an embrace of the norm to respond to the voice, with all of its consequences for our behavior. One might think that there is a performative contradiction here. After all, has Cavarero not criticized the history of philosophy for ignoring a norm that she herself argues for in a fashion that cannot be called anything other than philosophical? Although sometimes Cavarero can be sweeping in her rejection of the philosophical project, I believe that there is no necessary contradiction here. A sympathetic reading of Cavarero would see her not as claiming that philosophy must necessarily exclude a recognition of the voice. That would indeed present a problem for her. Rather, we should read her critiques of philosophy as historical ones. Philosophy, in her view, has developed a set of norms that have divorced it from the voice and thus from the individuality of humans. Her project, then, is not so much to turn away from philosophy in order to recover the voice but instead to reorient philosophy by arguing for a normative approach to which its history has blinded it. In this way, her project can be seen on analogy with Foucault’s genealogy, which does not constitute a rejection of history but rather an insistence that in approaching history in a genealogical rather than progressivist fashion, we can see aspects of ourselves more clearly than the latter approach would allow. I have insisted throughout this paper, and have tried to exemplify with the brief treatment of Cavarero’s work, that normativity is deeply effable, that is, that it is bound to linguistically articulable standards. I have not insisted that it is reducible to those standards, and have, in fact, argued that it is not. However, I would like to push against any attempt to read thinkers like Levinas, Derrida, and Cavarero as showing us that normativity is betrayed when it is given semantic form, for the reasons I have canvassed. To be sure, I have not provided an argument that all normativity must be effable in the ways I have described. Instead, I have offered several examples and a framework to account for them. But I believe that it would be difficult, indeed

The Effability of the Normative  211 perhaps impossible, to find a normative realm in which this framework does not hold. In closing, I would like to call attention to the fact that not only does it seem to me to be a feature of normativity to be bound, although again not reducible, to linguistic articulation. It is a very good thing that it is so bound. Any normativity that cannot be assessed at the linguistic level I have described would be a normativity immune to criticism. It would lie outside our ability to reflect upon it, ask whether its norms are appropriate or tolerable, and act according to our own reflection. Such norms would be dogmatic, not in the sense that they would be undefended beliefs, but rather in the more pernicious sense that they would be undefended orientations of our behavior. Anyone who is not entirely content with the normative structure of the current world should find that prospect a frightening one. So, against anyone who would seek to condemn philosophical reflection for reducing normativity to the totalitarianism of the effable, I would argue that it is precisely such philosophical reflection that allows us a critical stance. Not only is effability an inescapable feature of our normative landscape; we are lucky that it is so.

Notes 1 Gutting (2013). 2 James (2012, p. 4). 3 See Wittgenstein (1958, sections 201–240). 4 Stanley (2011). 5 In Otherwise Than Being, Levinas explicitly lays out the idea that normative judgment—the said—is founded on a primordial saying that is grounded in an ineffable proximity to the Other. In his discussion of justice and the “third party” that is outside the dual ethical relationship of “self” and “other,” justice and representation arise. “In the comparison of the incomparable there would be the latent birth of representation, logos, consciousness, work, the neutral notion being.” Levinas (1981, p. 158). But this representation, which is where justice and equality reside, is founded on the inegalitarian infinite responsibility of each for the other. “The equality of all is borne by my inequality, the surplus of my duties over my rights.” (p. 159) 6 Derrida (1973, p. 130). 7 Foucault (1977). 8 Wittgenstein (1958, Section 217). 9 Wittgenstein (1969, Section 253). 10 May (2001). 11 Cavarero (2005, p. 175). 12 Ibid., p. 176. 13 Ibid., p. 207.

Works Cited Cavarero, Adriana. 2005. For More Than One Voice: Toward a Philosophy of Vocal Expression, translated by Paul A. Kottman. Stanford: Stanford University Press. Derrida, Jacques. 1973. Différance. in Speech and Phenomena and Other Essays on Husserl’s Theory of Signs, translated by David Allison. Evanston, IL: Northwestern University Press.

212  Todd May Foucault, Michel, 1977. Discipline and Punish: The Birth of the Prison, translated by Alan Sheridan. New York, NY: Random House. Gutting, Gary. 2013. Thinking the Impossible: French Philosophy Since 1960. Oxford: Oxford University Press. James, Ian. 2012. The New French Philosophy. Cambridge: Polity Press. Levinas, Emmanuel. 1981. Otherwise Than Being, Or Beyond Essence, translated by Alphonso Lingis. The Hague: Martinus Nijhoff. May, Todd. 2001. Our Practices, Ourselves. University Park, PA: Penn State Press. Stanley, Jason. 2011. Know How. Oxford: Oxford University Press. Wittgenstein, Ludwig. 1969. On Certainty, translated by Denis Paul and G. E. M. Anscombe. New York, NY: Harper Torchbooks. Wittgenstein, Ludwig. 1958. Philosophical Investigations, Third Edition, translated by G. E. M. Anscombe. New York, NY: Palgrave Macmillan.

Part III

The Significance of Indeterminacy for Hermeneutics and Aesthetics

10 Indeterminacy, Gadamer, and Jazz Bruce E. Benson

Without indeterminacy, jazz is not only impossible but also unthinkable. The very act of improvisation requires that at least something has not already been determined. Yet jazz is equally dependent upon determinacy. Without determinacy, jazz is not only impossible but also unthinkable. It is the interplay between determinacy and indeterminacy that makes jazz possible. In what follows, I wish to consider how determinacy and indeterminacy relate. My thesis is that they are entirely dependent upon one another in everyday life and logically dependent upon one another to the point of definition. To make sense of this claim, I turn to the work of Hans-Georg Gadamer, particularly his major work Truth and Method. Gadamer provides a way of thinking about indeterminacy that allows us to see just how undetermined things really are without causing us to fall into the nihilistic abyss. If the tendency in modern philosophy (here I use this phrase in its technical sense) has been toward increasing and ever more precise determination, the postmodern tendency has been generally toward the questioning of the very possibility of precise determination. There is something to be said for each of these approaches, so I will not be preferring one to the other. As will become clear, if either determinacy or indeterminacy, certainty or uncertainty, clarity or opacity are privileged, we arrive at a view that fails to provide a properly phenomenological sense of the world. By “properly phenomenological” I simply mean “that which comports with our experience.” I use the example of jazz since it provides a concrete example of this relationship between determinacy and indeterminacy, which means that it helps us to think more deeply about their relationship. At the heart of this issue is the notion of “spontaneity.” What is it to be spontaneous?

The Relation Between Determinacy and Indeterminacy Indeterminacy is not simply a fundamental feature of human experience, it is a fundamental feature of the universe—at least the universe that we know. Perhaps there could be some other universe that is fully determinate. If so, life in such a universe would have to be very much unlike life as we know it. For living beings are able to be precisely because they are able to become

216  Bruce E. Benson what they are not. Aristotle is perhaps best remembered for making such a point, though Heraclitus had likewise reminded us that change is fundamental to all there is. Of course, both had to have recourse to some other “thing” that does not change—for Heraclitus, this is the logos; for Aristotle, the “Unmoved Mover”—to explain how change takes place. While we might think that we have now moved beyond such primitive thinking, the very fact that we need to have recourse to the notion of, for instance, “personality” to make sense of the dynamic nature of human beings shows that we really haven’t. It is hard for us to “think” about a world in which everything really is in motion. In other words, things can’t boil down to sheer spontaneity. Of course, our everyday thinking is way behind the advances in theoretical thinking in such areas as particle physics and quantum mechanics. We can only ascent to such theories as long as we are not forced to see them working out in practice. We may agree that time ultimately is relative, but we cannot really think this way in practice. Consider how the Hebrew Bible opens: we are given a story of Creation. As it turns out, this story is more or less the same for other ancient Mesopotamian societies. For our purposes here, whether this is meant as a version of “on this day in history, this is what really happened” or it is simply a founding myth, is completely unimportant. The reader is told: “In the beginning when God created the heavens and the earth, the earth was a formless void and darkness covered the face of the deep” (Gen 1:1–2a). This strange term “formless void” is an attempt to render the phrase “tohu va vohu,” itself a strange phrase. While the standard doctrine in orthodox Christianity is that of creatio ex nihilo, this is not the formula given here. Instead, we are left with a curious mixture of determinacy and indeterminacy. That there is something at all is the determinacy part; that it is formless is the indeterminacy part. But it is important to note that, in one important sense, the story moves toward ever greater determinacy. Out of the void comes light, which is promptly separated from darkness. From formlessness comes the distinction of earth and sky, and then the distinction of earth and sea. The move is from indeterminacy to determinacy. There are two points that such a story raises. First, we tend to like stories in which indeterminacy is replaced by determinacy. Isn’t a story of increasing determinacy much more settling and comforting than a story in which indeterminacy is replaced by more indeterminacy? Here we can turn to a point made by Friedrich Nietzsche regarding stories we tell to ourselves. Nietzsche makes the rather simple but profound point that the lies we tell are principally to ourselves. As he puts it, “lying to others is relatively an exception.”1 What Nietzsche is saying is that we like narratives about ourselves in which we are the good people and those other people are the bad people. We avoid narratives in which we see ourselves going down a wrong path. But, more important for our point here, we want narratives that are not overly complicated or indeterminate. Narratives in which we sometimes make good choices and sometimes make bad choices—i.e., those in which

Indeterminacy, Gadamer, and Jazz  217 we are inconsistent and unreliable—are ones we tend to avoid. This inconsistency is why we feel a need to tell ourselves lies. Our actual selves are rather messy and indeterminate. Indeed, much of our everyday practice is murky and riddled with either simple contradictions (we espouse X but do Y), or we do actions that are at odds with one another. We speak of someone as having “integrity,” which means that the various parts of the person fit perfectly together to form an integrated whole. However, integrity is always a goal to be reached rather than something one has already attained. Aristotle reminds us that even the “phronimos”—the one who lives a life of practical virtue (phronesis)—is still always on the way to becoming more perfect in this respect. Yet Aristotle likewise has a problem, which he terms “incontinence.” In such a case, a seemingly moral person commits a heinous act that seems to come from out of the blue. For Aristotle, the only explanation for such an action is that the person was never really moral in the first place. However, the problem with Aristotle’s position is that the actual phenomenon of being a human being doesn’t go nearly as neatly as he thinks it does. Someone can be both moral in impressive ways and yet be immoral (or, more kindly put, lacking) in significant ways—at the same time. Literary and religious stories are full of people who are “heroes” but sometimes even tragically flawed. Odysseus has a very mixed character, one that can elicit almost equal amounts of praise and blame. The Christian tradition has the notion of saints, who would seem to have achieved moral perfection. But, if one looks closely at the lives of even the most celebrated saints, one discovers that—to a great extent—they wouldn’t be all that nice to have around on a daily basis. They may have done impressive things, but they often are ambiguous. Tertullian, Jerome, Cyril of Alexandria—all of them were deeply flawed. And yet they are celebrated as saints—a determinacy that is intricately interlaced with indeterminacy. This leads us to a second point. While it might at first seem that the creation story in Genesis simply moves from indeterminacy toward ever greater determinacy, the actual telos is much more complicated, and more interesting. For, while there are determinations—such as the bringing into being of particular creatures—what those creatures in turn go on to do is only partially determined. To be sure, they are fruitful and multiply. Yet, the more they multiply, the more variations and thus indeterminacies become possible. The reality of the world is that indeterminacy is never replaced simply with determinacy. Instead, it is replaced with yet another mixture of indeterminacy and determinacy. And this is always the case. If we turn back to the example of human persons, we see that they start out with certain determinations but many indeterminations. One is born in a particular place, but one may end up moving to another country or even a series of different geographical points. One’s parents are two specific beings, though such a story used to be somewhat simpler than it is now in an age in which one’s parents may not be male and female (even though one’s biological parents are—so far). To grow up is to have more determinations. One attends

218  Bruce E. Benson school and has particular teachers, particular friends (and enemies), and particular subjects of study. However, each of these “particularities” opens up a whole new world of “indeterminacies.” True, once you’ve finished the first grade, you can’t go back again (well, unless they require you to do so). But moving on closes certain doors and opens up other ones. It might be thought that what I’ve said in the last few paragraphs is meant as a kind of negative commentary on human failings. Such is not my point at all. Rather, I am simply using the actual existence of human persons to make it clear that indeterminacy and determinacy are not really all that separable. One can be determined in some ways and not others. Yet even that observation forgets the existence of time. To be historical beings means that we change over time. Such change can move toward more determinacy. Yet it can just as well be toward more indeterminacy. Or, better put, even the move toward determinacy likewise opens up possibilities for indeterminacy, and the other way around. For example, one of the benefits (or is this a drawback?) of being a philosopher is that, over time, issues that seemed so straightforward in one’s youth often appear much more complicated than originally thought. The movement is nothing as simple as, say, moving from the view that morality is fixed to a view in which everything is permitted. Rather, how one even thinks about “right and wrong”—or the correspondence theory of truth or the nature of human freedom—becomes much more nuanced. One can, for instance, come to see that human beings are much less free than one originally thought without simply giving up the notion of freedom. Here it is helpful to make a connection to the OED’s second definition of indeterminacy. The phrase given is “want of determinateness or definiteness” and I assume here that “want” is really just another word for “lack.”2 Yet, even then, there is the question of how we are to understand this “lack.” On the one hand, it can simply be interpreted as something that is not there. On the other hand, this “want” could also be taken to be something like a “desire” for determinacy. Put differently, there is no such a thing as “lacking” without the perspective of a subject, in either the simple, seemingly straightforward sense or in the prescriptive sense. It may be that God thinks that something is lacking. More likely and immediately, though, it is we who see a lack. Yet, herein lies the problem. While there are certainly ways in which indetermination is negative (inconsistent behavior toward one’s friend, for instance), indetermination is absolutely required if there is to be development and freedom. In other words, one must not see this as simply going one way and not the other. But we can make this point even stronger. Note that simply to make the point of indeterminacy there must be some determinacy. I cannot indeterminately make any point about indeterminacy. Nor can indeterminacy be thought indeterminately. This interplay is what I mean in claiming that each is dependent upon the other. Put in the strongest possible terms: anything like “pure indeterminacy” is unthinkable and impossible; anything like “pure determinacy” is

Indeterminacy, Gadamer, and Jazz  219 probably just as unthinkable, but it would also not characterize anything as we know it. In Christian theology, there are debates between those who favor more kataphatic conceptions of God, in which God is characterized in clear terms, and those who favor more apophatic conceptions of God, in which clarity regarding God is put into question. Yet these are always going to be emphases rather than a simple opposition. For human beings, freedom is precisely this varying mixture of indeterminacy and determinacy. And this is true for both negative freedom—the freedom from constraints—and positive freedom—the freedom to be able to do something. So far, my discussion of indeterminacy might itself be characterized as in “want” of determinacy in both of the senses we considered above. Let me move, then, to a more concrete kind of example, in this case the interpretation of texts.

Gadamer and the Sense of Taste One of the many merits of Truth and Method is that Gadamer attempts to give both determinacy and indeterminacy their proper due. Gadamer writes as a phenomenologist and, ultimately, a phenomenology succeeds or fails on the basis of whether it adequately describes the phenomenon (or phenomena). The only way to challenge a phenomenology is to provide a counterphenomenology, which boils down to saying: “no, it doesn’t go like that; it goes like this instead.” Perhaps the most important reason why Gadamer is so successful is that he is attentive to the multiple complexities and nuances of both texts and readers. Yet it is likewise important that Gadamer begins in a place that one might not expect. He turns to classic notions in what he terms the “humanist tradition.” Although his point only becomes fully clear as the text progresses, already from the start Gadamer reminds us that these notions are more fundamental than questions of science and method.3 Why? Because these notions are absolutely necessary both to think about questions of method and to employ scientific methods. Gadamer begins with Hermann Helmholtz’s distinction between logical induction and artistic-instinctive induction. Already, the difference should be clear: whereas the former can be reduced to reason and certainty, the latter cannot. Yet this in no way invalidates it, for it is not at all a matter of individual caprice. Not surprisingly, Gadamer points out that Helmholtz makes a connection between this second sort of induction and “tact.” Gadamer writes: “By ‘tact’ we understand a special sensitivity and sensitiveness to situations and how to behave in them, for which knowledge in general principles does not suffice.”4 At this point in English usage, the term tact refers almost exclusively to social and political relations. We talk about someone showing tact in, for instance, dealing with an awkward social situation (such as when a relative has had a bit too much to drink at a wedding reception) or dealing with the leader of a foreign government who makes unreasonable

220  Bruce E. Benson threats. Knowing how to approach such persons and diffuse the situation often requires an exquisitely tuned sense of tactfulness. Of course, this is a kind of knowledge. While we tend to reserve the term “knowledge” for things that are more concrete or logically demonstrable, knowing how to “get along” with people is a crucial and basic kind of knowledge, a point that has become all the more obvious as we come to understand how people on the autism scale can miss social cues and thus have trouble interacting with other people. In effect, Gadamer provides a mini genealogy of the kind of knowing that requires being a particular sort of person. To exercise tact, one must actually be tactful. And this leads Gadamer back to the notion of judgment. “The word ‘judgment’ was introduced in the eighteenth century in order to convey the concept of judicium, which was considered to be a basic intellectual virtue.” As Gadamer puts it, “the difference between a fool and a sensible man is that the former lacks judgment—i.e., he is not able to subsume correctly and hence cannot apply correctly what he has learned and knows.”5 Yet what is meant by “subsume” in such a case? In essence, judgment is the ability to know how to apply a universal (such as “do not steal”) to a particular (the person you note at an electronics store who appears to be attempting to shoplift). What Immanuel Kant realized is that the faculty of judgment is such that it cannot “demonstrate” that it is correct. The only way to do so would be to appeal to another faculty (and this would lead to an appeal to yet another faculty, ad infinitum). Gadamer concludes that judgment “cannot be taught in the abstract but only practiced from case to case, and is therefore more an ability like the senses. It is something that cannot be learned, because no demonstration from concepts can guide the application of rules.”6 To make this point clearer, we could (following Kant) point out that, if one needed a rule in order to help one follow a rule, the result would likewise be an infinite regress. At this point, it should be clear that a judgment is a determinacy that is still somewhat indeterminate. A judgment can never be “justified” by a logical argument. If I believe that I have acted aright, the best I can do to convince someone else of this point is to lay out the specific circumstances, the persons involved, the intended benefits to be derived, and much more. Should I fail to convince this person that I acted in the most reasonable way possible—that is, that I acted justly—the best I can do is provide an even more-nuanced account of what factors were part of the decision and how they were weighed. Fortunately, much of the time, most other people—when presented with the same set of data—will agree that this was “the right thing to do.” Should that fail, however, the best I can do is say, “I made a judgment call.” Whether it is the coach who decides to remove one player from the game and send in another or the CEO who fires a particular employee, the post-game analysis or the discussion at the board meeting may vindicate such an action. Yet, if that’s not the case, there is no “higher” authority to which one could submit the issue for resolution (unless we mean something

Indeterminacy, Gadamer, and Jazz  221 like a committee appointed to investigate the matter that might itself have to fall back on “we made a judgment call”). How should we think about such instances? To begin, note that these kinds of “judgment calls” are not some unusual thing: we make them every day, and often about things that are fundamentally important (such as what college we attend or what field of employment we enter or whom we choose as a life partner). In many cases, such decisions are fundamental in shaping our future. Yet the problem is that none of the instances I named are such that they could be decided in any kind of “formal” way: there is no “equation” one can do to choose one’s major in college. In such a case, we might well do a “cost/benefits” analysis, but this is never fully conclusive in the way that an answer to a math problem is. Following Kant, yet also going historically behind Kant, Gadamer introduces the notion of taste. Admittedly, this might seem like a strange connection. Certainly, in English the term taste generally is restricted to matters of aesthetics or food or fashion preferences. Thus, when we say to someone, “You have good taste,” we generally think we are more or less saying, “Your taste is the same as mine.” Indeed, taste would seem to be one of the most “relativistic” kinds of things there are. You like Picasso; I like the impressionists. You like vacations in Italy; I prefer them in Australia. Yet, Gadamer follows Balthasar Gracian by going back to the notion of taste, “this most animal and most inward of our senses,” and points out that it “still contains the beginnings of the intellectual differentiations we make in judging things.” It is taste that we use to make fundamental judgments. But, even more important, “the concept of taste undoubtedly implies a mode of knowing. The mark of good taste is being able to stand back from ourselves and our private preferences. Thus taste, in its essential nature, is not private but a social phenomenon of the first order.”7 As I mentioned before, we often think of taste in regard to aesthetic phenomena and assume that taste is no more than merely “I like X.” But taste cannot be so narrowly defined. There truly is something like “good taste,” which is why aesthetic relativism is wrong. I may not “like” Picasso in the sense of wanting to own one of his paintings and hang it in the living room. But, if I have any sense, I will give Picasso his due and recognize that he is one of the most important painters of the twentieth century. Further, taste can be cultivated: one can move from simply liking white zinfandel to developing a much more complex palate. So we are left in a most curious situation. As Gadamer says, “[G]ood taste is always sure of its judgment—i.e., it is essentially sure taste, an acceptance and rejection that involves no hesitation, no surreptitious glances at others, no searching for reasons.”8 In this sense, taste is like one of the five senses. If I believe I see a book on the table in front of me, the best I can do is try another sense (like the sense of touch) and reach out to feel if it is there. But such an example is only really for philosophy papers! Any sensible person would simply accept what she sees. Only highly unusual circumstances

222  Bruce E. Benson would require anything else. So, taste is determinate, not indeterminate. And, yet, the problem would seem to be that, if it is like a sense, then it cannot give any further justification of itself. In that sense, it would seem to be indeterminate. Yet, here the problem seems to be more one of how we talk about taste then of how it operates. Gadamer is right to point out that the opposite of good taste is not bad taste but no taste. The person who doesn’t “get it” is the one without any taste. Further, as we will see in a moment, the person without taste is not properly part of the community. Instead of thinking of taste as an individual thing, we need to see it as a fundamentally social phenomenon, such that having good taste is all about being part of a community.

The Indeterminacy of Jazz Improvisation As a distinctly musical concept, “indeterminacy” arose from the work of such composers as John Cage. In the section titled “Indeterminacy” in his book Silence, Cage cites Johann Sebastian Bach’s The Art of the Fugue as an example of an “indeterminate” work in the following sense: “The function of the performer, in the case of The Art of the Fugue, is comparable to that of someone filling in color where outlines are given.” Cage goes on to point out that this may be done “in an organized way” or in a more unorganized way, such as the performer “arbitrarily, feeling his way, following the dictates of his ego.”9 Cage considers his music to be “experimental” in nature and writes that “an experimental action is one the outcome of which is not foreseen.”10 The emphasis, then, is on “spontaneity” and “uniqueness.” One doesn’t have to read much literature on jazz improvisation to know that the emphasis is usually on precisely the spontaneous nature of jazz. Perhaps Cage was aware that the term improvisation comes from the Latin term improvisus, meaning “unforeseen,” so that what he is describing as experimental action sounds much like improvisation. Yet even the venerable Oxford English Dictionary defines improvisation as “the act of improvising or composing extemporare” and defines “improvise” as “to compose (verse, music, etc.) on the spur of a moment.”11 Not surprisingly, such a definition reflects the views of most people and, up until recently, even of musicologists (that is, people who should know better). However, this view is itself indebted to the ways in which “composition” and “performance” have been defined, though these definitions are themselves of relatively recent origin and only represent performance practice of classical music for roughly the last two centuries. In contrast, one need only turn back to Baroque music to find a very different conception of composing and performing. David Fuller makes the point that “a large part of the music of the whole era was sketched rather than fully realized, and the performer had something of the responsibility of a child with a colouring book, to turn these sketches into rounded artworks. . . . The closest modern parallel to the gap between Baroque notation

Indeterminacy, Gadamer, and Jazz  223 and the sounding produced is to be found in jazz.” Fuller then goes on to say that these performances varied “from one group to the next, one day to the next, one neighbourhood to the next.”12 It is highly instructive that he titles his chapter on Baroque performance “The Composer as Performer,” because the compositional process was not limited to the composer but extended in such a way to the performer that no clear distinction could be drawn between the two roles. Or one could see the performer as improviser, given that no performance would have been possible without improvisation. At stake here, of course, is the very nature of indeterminacy. Let me work this problem out by considering the phenomenology of music sketched by Roman Ingarden, who wants to make a clear-cut distinction between 1) the work itself, 2) the score, and 3) the various performances of the work. Needless to say, we do make a distinction between these entities. The question, though, is how strongly that distinction can be held. No one would mistake a given performance for “the work itself.” Performances of the “same” work can happen at the same time (say, in Paris and Berlin) or over the course of centuries. In this important sense, we can say that performances cannot be repeated. Even in the case of a recording, to play it once again is to set it into the context of this time and place rather than the original context. One might be more tempted, though, to say that the work itself is the same as the score. After all, the score provides the contours for the performance. It all depends, though, on which score one has in mind. On the one hand, Baroque music provides relatively few contours for the performer, who is given a great role in determining how the music will sound. On the other hand, composers in the twentieth century often sought to determine the contours of their pieces as concretely as possible and demand that performers adhere to these “instructions.” Yet the problem with musical notation is that it is impossible for the composer to establish all the features that go into a performance. Ralph Vaughan Williams, who was both a composer and a conductor, describes the score as “most clumsy and ill-devised” and, as a result, “has about as much to do with music as a time table has to do with a railway journey.”13 A score does not tell us much about the musical experience, since it is so underdetermined by the score. There is so much to a musical performance that needs to be added in order for there even to be a performance. Yet, even if there really were true determinacy, that does not mean that what is determined is precise. Consider what E. D. Hirsch, who famously insisted that a text means what the author intends it to mean, says. “Determinacy does not mean definiteness or precision. Undoubtedly, most verbal meanings are imprecise and ambiguous, and to call them such is to acknowledge their indeterminacy: they are what they are—namely ambiguous and imprecise—and they are not univocal and precise.”14 In other words, determinacy still entails indeterminacy. There is no point in writing something— a book, a piece of music—at which everything becomes precise. At one point, Ingarden insists that all aspects of a musical composition “have been

224  Bruce E. Benson conceived by the composer as fully defined and fixed.”15 Yet later he goes on to say: “Strictly speaking, before the performance of his work, even the composer himself does not know the profile in all its qualifications; at best he imagines it more or less precisely and at times he may merely be guessing at it.”16 Of course, even after the performance of the work, the composer has only heard one particular interpretation of it, not all possible interpretations. Further, it is hard to imagine that the composer hears the performance and thinks, “Yes, that’s what I intended all along,” regarding all aspects of the performance. It is much more likely that the composer will find some aspects pleasing and others less so. In any case, all that a performance does is provide the exact contours for that performance; it hardly provides them for all other performances. Given these problems, Ingarden comes up with the idea of what he calls “unbestimmtheitssellen” or “places of indeterminacy.” The idea here is that the composer knows certain things that she intends the score to convey but not others. In practice, this means that even fully notated music is still indeterminate. Only in a given performance are those contours made determinate; yet each performance determines them in different ways (and simply the attempt to repeat a given performance already shows this to be the case—there is no pure repetition; try as one might to make it sound exactly like the other proves impossible). The problem of the indeterminacy even in determinacy is something that jazz musicians actually celebrate. In order to see this, we need to turn to an aspect in jazz that is normally greatly overemphasized: spontaneity. If classical music would seem to be about closing off possibilities to performers (“play these notes this way”), jazz would seem to be all about opening them up. Yet, if indeterminacy lies at the very heart of determinacy, so determinacy lies at the heart of indeterminacy. If we go back to Cage’s conception of “coloring”—which he says is what the performer does with Bach’s The Art of the Fugue—we arrive at a much better way of thinking what transpires in jazz improvisation. The jazz composer/performer Carla Bley says this about her pieces: “I write pieces that are like drawings in a crayon book and the musicians color them themselves.”17 In this conception of composing, the composer provides something like an “outline” that performers then fill in during the performance. How this is done will depend on a number of factors. One is simply the kind of jazz one performs. One extreme would be “Dixieland” jazz. The kind of music performed at Preservation Hall in New Orleans is considerably determined simply in terms of its form. If one of the musicians were suddenly to take a solo, this would not only be frowned upon but also change the very form of music being performed. Within the conventions of Dixieland jazz, there is certainly room for the individual musician to improvise, but this space is relatively small. The other extreme is what we call “free jazz.” The artist/musician Sam Rivers claims “there’s nothing I can do wrong [sic], nothing.”18 While it is true that free jazz has a great deal of indeterminacy, if it were simply pure indeterminacy there could be no improvisation. Part

Indeterminacy, Gadamer, and Jazz  225 of this is because even free jazz works with certain conventions. I find it interesting, for instance, that free jazz is almost always dissonant in nature. If one considers various free improvisations, it becomes clear that there are still some structures at work. On her NPR program “Piano Jazz,” Marian McPartland regularly improvised in a sort of way (not following a particular tune or even a distinct pattern). Yet, her improvisations were never completely surprising or spontaneous. The dampening down of a player’s “spontaneity” is all the more true when one is improvising with other people. A major part of improvising (in music, to be sure, but even more in theater) is listening to other people and adjusting one’s playing accordingly. One listens to what the others are playing and works to “fit in” with the emerging dialogue. To do so requires exactly what we have seen so far—tact, taste, judgment. It should be stressed here that these aspects are both aesthetic and ethical. If one has a solo, there is usually something like an agreed upon number of times one is allowed to repeat a chorus. If a player has a really “hot” evening—everything is really cooking—one’s fellow players are going to be more generous about how long they allow the player to keep soloing. After all, it seems a shame to cut off something that’s really getting going in a good way. And, yet, how much freedom that player is going to be allowed—even on a great night—is both an aesthetic and ethical concern. As an aesthetic consideration, an improvisation on, say, Night and Day needs to have a balance between players and instruments. Perhaps it might be acceptable if a particular person in the group has more of a presence on a particular tune; but the balance is quickly ruined when one member continually dominates. Moreover, that aesthetic ruination is very closely connected to an ethical problem. If I choose to (selfishly) dominate in playing, that shows a profound lack of respect for the people with whom I’m playing. One lacks both taste and tact. One is guilty of poor judgment—aesthetically and ethically. Of course, even this question of balance has much to do with the ground rules for a given group. It’s normally the case that, in a piano trio, one hears much more of the piano. However, the pianist Bill Evans—particularly in his trio that included Scott LeFaro (bass) and Paul Motian (drums), which gave rise to the famous Village Vanguard recordings—deliberately gave LeFaro a great deal of space to improvise, seeing the relationship as more of a partnership. How exactly a particular group will see itself as working together is something that is both established in practice and in performance. One aspect that we have in effect presumed but on which we have not focused is something central to Gadamer’s discussion. It is the notion of Bildung. Gadamer quotes Wilhelm von Humboldt as saying “when in our language [German] we say Bildung, we mean something higher and more inward, namely the disposition of mind which, from the knowledge and feeling of the total intellectual and moral endeavor, flows harmoniously into sensibility and character.”19 Traditionally, Bildung has been the term for education and culture, but here the emphasis is on how one becomes

226  Bruce E. Benson a person of taste and discernment. Gadamer makes the connection with Formierung and formation: both have to do with how one develops as a person. Jazz is entirely dependent upon formation. One needs to know the chords and the scales, and how to take a solo. To play jazz is to enter into a tradition that is both determinate and indeterminate, since the tradition itself is constantly being formed—even by the particular improvisations currently being played. Without becoming gebildet in the ways of jazz, one would never be able to perform with others, and even not be able to perform as a solo act very well, either. To be gebildet, one must know the history of jazz. One must have heard hundreds of hours of performances. One must have had to try—and fail. Indeed, one of the greatest dangers for the jazz improviser is to fall back on the determinate improvisations from past “greats.” Jazz finds its boundaries in constantly pushing beyond them. If there had not been the desire to escape from sheer determinacy, we would all be playing Dixieland jazz. In my own writing on jazz, the metaphor of a conversation has been fruitful, not simply in understanding jazz but also in understanding the play of determinacy and indeterminacy. Gadamer turns to the phenomenon of conversation much later in Truth and Method and writes the following: We say that we “conduct” a conversation, but the more genuine a conversation is, the less its conduct lies within the will of either partner. Thus, a genuine conversation is never the one that we wanted to conduct. Rather, it is generally more correct to say that we fall into conversation, or even that we become involved in it. The way one word follows another, with the conversation taking its own twists and reaching its own conclusion, may well be conducted in some way, but the partners conversing are far less the leaders of it than the lead.20 Conversations always have some determinate features. Without such features, there could be no conversation. Yet, if those features were fully determinate, there would also be no conversation. It is this interplay of determination and indetermination that allows for a conversation to develop. In conversation, one does not want to dominate—unless one simply wishes to hear oneself talk. Instead, all parties want (or should want) an understanding to emerge that would not have been possible to think before the interaction of conversation. When Gadamer goes on to say that “understanding or its failure is like an event that happens to us,” he gets at the way conversation really works out. It is this ever-shifting balance of determinacy and indeterminacy that allows for something to emerge. The challenge is allowing a conversation to unfold, and to listen.

Notes 1 Nietzsche (1954, §55). 2 The Oxford English Dictionary (2009, “indeterminacy”).

Indeterminacy, Gadamer, and Jazz  227 3 Gadamer makes explicit reference here to the Geisteswissenschaften, a term that imperfectly translates into the “human sciences.” But, ultimately, his point is that any kind of “scientific method” (for either the social sciences or the hard sciences) depends on these more basic aspects of human experience. 4 Gadamer (1989, p. 16). 5 Ibid., p. 30. 6 Ibid., p. 31. 7 Ibid., p. 35. 8 Ibid., p. 36. 9 Cage (1961, p. 35). 10 Ibid., p. 39. 11 The Oxford English Dictionary (2009, “improvisation” and “improvise”). 12 Fuller (1989, pp. 117–118). 13 Williams (1987, p. 124). 14 Hirsch (1967, p. 44). 15 Ingarden (1986, p. 116). 16 Ibid., p. 148. 17 I first discovered this quote in an airline magazine. Since I couldn’t trace its source, I ended up writing to Ms. Bley. She said that she didn’t remember saying it, but it sounded like something she would say. So she gave me permission to use it as an epigraph in my book The Improvisation of Musical Dialogue: A Phenomenology of Music (Cambridge: Cambridge University Press, 2003). 18 NPR Weekend Edition (03–28–1998). 19 Gadamer (1989, pp. 10–11). 20 Ibid., p. 383.

Works Cited Cage, John. 1961. Silence: Lectures and Writings. Middletown, CT: Wesleyan University Press. Fuller, David. 1989. “The Performer as Composer,” in Performance Practice, Vol. II, edited by Howard Mayer Brown and Stanley Sadie. Houndmills, UK: Palgrave Macmillan. Gadamer, Hans-Georg. 1989. Truth and Method, 2nd ed, translated by Joel Weinsheimer and Donald G. Marshall. New York, NY: Crossroad. Hirsch, Eric D. 1967. Validity in Interpretation. New Haven, CT: Yale University Press. Ingarden, Roman. 1986. The Work of Music and the Problem of Its Identity, edited by Jean G. Harrell, translated by Adam Czerniawski. Berkley: University of California Press. Nietzsche, Friedrich. 1954. “The Anti-Christ(ian),” in The Portable Nietzsche, edited and translated by Walter Kaufmann. New York, NY: Viking Press. NPR Weekend Edition. 1998. “Modern Jazz Pioneer Sam Rivers Profiled.” air date: 03–28–1998. The Oxford English Dictionary. 2009. 2nd ed. Oxford: Oxford University Press. Williams, Ralph Vaughan. 1987. The Letter and the Spirit. in National Music and Other Essays. Oxford: Oxford University Press.

11 Hermeneutic Priority and Phenomenological Indeterminacy of Questioning Nathan Eric Dickman

Through our cultural obsession with answers, we seem to be careening toward a “disinformation” age. Might shifting more emphasis to the significance of genuine questioning steady this? In this chapter, I will explore two questions about questioning itself. What significance does indeterminacy have within questioning? And, how does questioning make it possible to understand meanings? I will address these through the framework of hermeneutic phenomenology with the following procedure.1 First, I specify genuine questioning and isolate semantic indeterminacy as the primary foci at stake. Second, I develop a hermeneutic of questioning to explain that only shared questioning makes possible consideration of meanings as meanings. Third, I develop a phenomenology of questioning through an intentional analysis of both its noetic (acting) and its noematic (content) poles. The significance of indeterminacy here concerns the predicative possibilities suspended about a subject in question. I conclude by noting some areas for application.

Preliminary Clarifications My primary concern in this chapter is with what I call genuine questioning. Questioning is researched in many fields, such as classroom conversations,2 and even is heroicized as synonymous with philosophizing itself.3 A fundamental distinction recurs between what variably is called lower-order and higher-order questioning,4 guiding and grounding questioning,5 or naïve and phenomenological questioning.6 Genuine questioning belongs among the latter. To bring genuine questioning into relief by way of contrast, let us look at an example of erotetic logic—the logic of question and answer—in which an interrogative sentence is converted into the imperative mood. We can render “What is your name?” as “Tell me your name” (or more discretely, “Select one: My name is Muhammad. My name is Ruth. My name is . . . [etc.]”). Erotetic logic is formulated in terms of “epistemic imperatives,” disjunctive sets of propositions from which the answerer selects so that that the questioner comes to know which proposition is the answer.7 If

Hermeneutic Priority, Phenomenological Indeterminacy  229 getting the answerer’s name matters most, though, other tactics might work besides questioning—like a nametag glance. Although the name-seeker may be sincere, I restrict “genuine” questioning to instances that cannot be reduced to an alternative mood.8 Such questioning aims not at needed answers. If getting information is the sole function of questioning, then we might wonder with Emmanuel Levinas whether we are capable of a thought in which words have meaning.9 Needed answers make questioning deprivation driven and calculative. Genuine questioning is surplus driven, rooted in a desire of “one who lacks nothing,” a desire oriented by what Levinas calls “the order of the Good.”10 Can we question for the joy of it, tarrying solely with what questioning opens to us without feeling deprived of answers? Let us phenomenologically bracket our naïve attitude that subordinates questions to answers so we can explore this unique phenomenon of genuine questioning too often undervalued in our disinformation age. Let us also bracket our natural attitude about indeterminacy, our typical concern about whether “mind-independent” objects are indeterminate. I use indeterminacy primarily as synonymous with predicative possibilities. We determine a sentential subject when we discover or compose a fitting predicate for it.11 A sentential subject primed for a fitting predicate is in a state of indeterminacy. Consider “What is this?” The subject set in relief by the pronoun “this” is yet to be determined, so the question articulates a state of indeterminacy about it. The question suspends it in a network of predicative possibilities. Edmund Husserl calls the general horizon of interrelated predicative possibilities the Lebenswelt, or lifeworld.12 The lifeworld is neither some ineffable immediateness nor some emotional layer of experience. It designates, Paul Ricoeur writes, “the reservoir of meaning, the surplus of meaning of the life experience, which makes possible the [identifying] and [predicative] attitude.”13 Thus, Gail Weiss writes, “The indeterminacy of the horizon should be understood as operating not solely on the fringes of experience but at its center.”14 The lifeworld is a generative indeterminacy, or, as Weiss emphasizes, the “rich subsoil” of our “wild-flowering region” of experience. Let us turn to examine indeterminacy’s intersection with the hermeneutic priority of questioning.

Hermeneutic Priority and Indeterminacy Most philosophical studies of questions start with judgments or propositions and derive explanations of questions from them. They mention dialogue and shared inquiry only cursorily rather than analyze it as questioning’s fullest form.15 Husserl, for example, brackets out considerations of dialogue in his analysis, and focuses solely on the solitary ego’s experience of frustration expressed in questioning. For Husserl, like the erotetic model, questions express an experience of frustration over an uncertainty that strives

230  Nathan Eric Dickman for satisfaction in a decisive judgment.16 Unlike an objective judgment, questioning merely expresses a subjective feeling. Genuine questioning, on the contrary, does not merely express my experience like the interjection “ouch” expresses my subjective experience of a pain. Karl Schumann and Barry Smith urge that questioning is “tied intrinsically to its being addressed to some other person.”17 Just as reading is always reading to someone, so is questioning done with someone.18 In this section, I explain questioning’s “intersubjective” or—more precisely—its hermeneutical dimension, with an aim at identifying this dimension’s significant indeterminacy. In philosophical hermeneutics, to understand a question is to ask it, but to understand a meaning is to consider a judgment as a potential answer to a question.19 This principle articulates questioning’s hermeneutic priority, and we can use it to draw out genuine questioning’s intrinsic intersubjectivity. We need to get down first that understanding questions does not belong to the same order as understanding meanings. Questions are not answers, so they do not “mean” anything.20 Answers do, and do so predominantly in the form of judgments, complete thoughts with a subject and predicate. Questioning is necessary to catch judgments as meanings. However, we do not have to “mean” a judgment even if it answers our question. A judgment is a meaning because we can consider it as among numerous possible answers, but it takes appropriation to “mean” it. By contrast, the only way to understand a question—to “mean” it—is to ask it. As Hans-Georg Gadamer points out, there is no “tentative or potential attitude of questioning. . . . Even when a person says such and such question might arise, this is already a real questioning that simply masks itself, out of either caution or politeness.”21 We might decide whether to utter one publicly, but we cannot decide whether a question occurs to us. For a question to be what it is requires adoption of it. Getting caught up in questioning, though, is less an intentional activity and more a passivity.22 Questions occur to us, even “strike” us. While we might decide to command others with interrogative sentences, genuine questioning’s occurrent character—not its intentional character—distinguishes it as questioning. That questioning occurs to us exposes one interesting dimension of indeterminacy crucial for ethical sharing.23 When I hear or read another person asking a genuine question, I necessarily ask the question too if I understand it. When we both ask the question, whose question is it: the other’s or mine? When I hear you ask a question and I understand it (and thus ask it), it transforms into our question. You might initiate it, but here it is shared. For illustration: my students have difficulty grasping Aristotle’s theory of friendship when they do not share Aristotle’s questions to which his theory answers. What is the most complete kind of friendship? If the students do not suture their own questions about friendship with Aristotle’s, they are unable to consider Aristotle’s argument as a possible answer, and so his conclusion will not be understood as a meaning. If sharing is achieved, though, then whose question is it? The indeterminacy of “us” asking the question

Hermeneutic Priority, Phenomenological Indeterminacy  231 helps maintain orientation toward the real subject: it is neither you nor me, nor even us, but the subject matter. Like absorption in games, genuine questioning draws us in and fills us with the animating spirit of the subject.24 This indeterminacy, moreover, keeps genuine questioning ethical rather than exploitive. Some critics claim questioning subordinates the other to one’s domination.25 If questioning demands an answer, it seems oppressive. As Slavoj Žižek illustrates, the authoritarian says, “It is I who will ask the questions here!”26 While perhaps fitting for epistemic imperatives, genuine questioning circumvents this because the questioners’ positionalities are indeterminate. Genuine questioning is without deliberate domination. Since only answers count as meanings to be understood, the other is respected as a co-explorer and is not assimilated.27 What we understand is not the other, but what the other states as potential answers to our shared questioning. It allows us to listen to what another says, opening a space for the other’s unfolding.28 Genuine questioning proves not to be a privative need to subordinate others, but a desire for a surplus of possible meanings through shared exploration. Questioning’s hermeneutic priority is further evident in that it makes shared meanings or “fusion of horizons” possible. Consider Paul Tillich’s explanation for ineffective Christian evangelism: The difficulty with the highly developed religions of Asia . . . is not so much that they reject the Christian answer as answer, as that their [conception of] human nature is formed in such a way that they do not ask the question to which the Gospel gives the answer. To them the Christian answer is no answer because they have not asked the question to which Christianity is supposed to answer.29 The point is not about forcing others to ask our question. Refusing to ask another’s question—or refusing to learn their language—may be in one’s best interest, especially if the others are complicit with imperialism.30 The point is that an answer makes no sense abstracted from asking the question it addresses. Questioning animates the consideration of judgments as meanings.31 Judgments as mere propositions, by contrast, are abstracted from shared questioning and possible answers. Michel Meyer calls such abstraction “propositionalism” and claims it has distorted Western thought, noting its corrosive effect on “logical reasoning.” In formal logic, a proposition removed from questioning is called a “meaning.”32 An argument consists of a string of judgments—premise(s) and conclusion—where validity concerns solely the inferential relations among them.33 The propositionalist ideology enframes “reasoning” as tracking inferences, where trigger-happy fallacy identification is the mark of “critical thinking.” Introductions to reasoning teach us to analyze validity without understanding the propositions. This occludes the questions these judgments answer, and so they are not “meanings” in our sense. As a kind of “idle talk” analyzed by Martin Heidegger,

232  Nathan Eric Dickman propositionalism masquerades as understanding when it is not.34 Answers, however, are not inferred from questions even if we translate questions into imperatives about disjunctive sets of propositions. Not only does shared questioning make shared meaning possible, but it also makes judgments as meanings possible. Can we feel assured we ask the “same” question? Our purportedly private subjective perspectives seem to inhibit this. The suspension of our natural attitude toward questioning applies here. To share the same question is to share a language, the conventions for identifying subjects and enframing them with predicates. Grammatical conventions help us speak about the “same thing.”35 To share a question is to know the same language or to translate proficiently between languages. Subject matters stand out as topics in part because cultural traditions hand them down as important. By inheriting a language, moreover, we appropriate fields of predicative intelligibility or language-games that reflect our cultural Lebensform.36 To illustrate, let us ask the question, “What year is it?” Our responses depend on our forms-of-life. To answer “2018 CE” is to bring Christian coordinates to bear, or at least a secular lifeform bestowed by Christendom. Yet for Muslims, it is “1439 AH.” To ask what year it is in genuine questioning rather than to get information, we can reflect on what time itself is (the subject) and the politics of era-dating systems (the possible predicates). We can also reflect on relations between languages and truths, because both dates are true, yet both seem unable to be true in the same way at the same time—though it seems equally absurd to ask which is “truly true.” We have come to the edge of questioning’s hermeneutic priority and now will turn to phenomenology to elaborate on what questioners share when they question together. I want to emphasize that questioning just makes it possible to consider meanings. It does not force us to “mean” any of them. While understanding a question is to ask and appropriate it, we can understand a meaning without assuming responsibility for it. Considering which year or whose year it is does not entail adoption of one or the other era-dating systems. But to take up one option as an answer—especially as “the” answer—is to take responsibility for determination within the predicative possibilities of indeterminacy. In shared questioning we are already co-responsible on the way to fusing horizons.37 Let us examine indeterminacy’s intersection with the phenomenological openness of questioning.

Phenomenological Openness and Indeterminacy My emphasis on meanings may seem like I am promoting an idealistic phenomenology of questioning, as if we are after the isolated ego’s grasp of an abstract proposition. As Ricoeur asks, “That consciousness is outside itself, that it is towards meaning, before meaning is for it, and still more, before that consciousness is for-itself [as ego], is this not what the

Hermeneutic Priority, Phenomenological Indeterminacy  233 central discovery of phenomenology implies?”38 The existential subject does not have mastery over the meanings understood, let alone the meanings intended, because meanings inform self-consciousness before an ego “means” or intends them.39 Thus, genuine questioning’s intentionality is not analogous to the subject/object dualism of purported immediate sensation, where a subject has unmediated access to an object.40 Such a model for intentionality is misleading for openness to meanings. If we take “images,” for example, merely as weak subjective substitutes for “mind-independent” objects affecting our senses, we undermine how image-laden language enhances understanding.41 Not all images are residual sensations. Poetic images reverberate not because of things seen, but because of things said.42 We “see” some things in that we first read or hear—that is, understand— them.43 In this section, I develop a phenomenology of genuine questioning with an aim at identifying indeterminacy within it. I want to examine both its noetic (acting) pole and its noematic (act or content) pole.44 This difference appears in ordinary expressions, such as “asking a question” and “the question asked.” Merleau-Ponty puts forward questioning as the fundamental essence of all noetic activity.45 Noematic content, however, is not immanent to conscious activity.46 It is what one is conscious of. This content, however, is neither reified Platonic ideas nor abstract propositions in “mental” space, but actual objects under the epoché of our naïve attitude toward them.47 Bracketing brings into relief not objects in the “external” world, but the willed as willed, the judged as judged, and most central for us the questioned as questioned. What are the noetic and noematic dimensions of genuine questioning? The noetic or acting pole of genuine questioning is not an introspectively directed reflection on an ego’s experience, but an unreflective participation in the flow of everyday activity.48 Merleau-Ponty describes this openness as “an original manner of aiming at something,”49 and emphasizes questioning as our “ontological organ” and the “ultimate relation to Being.”50 It is more than sensory “aiming” because it includes contraction and enfolding so that we might receive, listen to, and consider potential answers. We can delineate this contraction via Merleau-Ponty’s notion of embodied dehiscence, the splitting of our body in two as simultaneously feeling and felt.51 This “split” embodies our fundamental openness so it is misleading to picture this as something torn or broken open. We naively believe we perceive “external” things with our senses, as if intentionality were a conduit transporting representations to our minds in a straight line. For MerleauPonty, we envelop things through the dehiscence of our bodies. This is why when we experience something, we say it is given, not through, but “in the flesh.”52 By means of dehiscence, the body gives itself to itself as a conscious network of potential or an “I can.”53 It is also oriented by anticipatory projections, or prejudices (literally, pre-judgments), which are effects of our conformity with inherited culture.54 Indeed, as Heidegger explores, our conformity is so thorough that we are first “the they” before wresting

234  Nathan Eric Dickman a self as an authentic individual.55 A self-conscious ego’s intentionality is subordinate to the primacy of embodied and historically conditioned features of experience.56 The historically conditioned dehiscent body shows us something interesting about genuine questioning. Without emphasis on embodiment, we may be tempted to abstract questioning, in Luce Irigaray’s words, “from its carnal taking root.”57 Questioning is as concrete as touch, where we grasp things in the folds of our flesh. Genuine questioning articulates the dehiscence of understanding so that meanings may be enveloped in its folds. How? Because questioning has the logical structure of suspending judgments (including prejudices), it maintains a distance from them while simultaneously being open to them as possible answers.58 Questioning does not eradicate prejudices, but puts judgments at risk by considering and testing them as possible answers.59 This is given “in the flesh” of spoken and written interrogative sentences. It is the locus where understanding “touches” itself, where questioning is the skin of receptivity and answers are the flesh of meanings. Meanings are sentences understood; sentences are meanings perceived.60 That answers disintegrate into contextless or questionless propositions may be natural because questioning does not “express” meaning.61 Questioning as questioning does not say anything; judgments do and are meanings insofar that they answer questions. In this way, questioning performs dehiscence, opening us to and enveloping possible meanings in the flesh. It might seem like sincerely wanting information is another noetic element of questioning, and this is emphasized across philosophical loyalties.62 Levinas criticizes this and argues instead that questions embody a different intentionality. If information is all we need, then we need not question another. Questioning does not have to aim at fulfillment in knowledge.63 “Must we not admit, on the contrary,” Levinas writes, “that the request and the prayer that cannot be dissimulated in [questioning] attest to a relation to the other person. . .? A relation delineated in the question, not just as any modality, but as in its originary one.”64 Questioning is about relationships more than about answers. Unlike an individual’s anxious need for answers, shared questioning is an opportune co-responsibility, a refusal of domination in openness to the subject matter and another.65 In genuine questioning, we do not know beforehand where we might end up or even if there must be some “final” answer.66 As Socrates describes it, we must follow the logos wherever it, like the wind, blows.67 The noematic or content pole of questioning seems to consist of the possible answers as a disjunctive set of propositions. Most theories of questioning begin here.68 Although we have subordinated the content to the intersubjective lifeform of questioning, it is still relevant for a more complete phenomenology of questioning. A key distinction that will help isolate the content is Gottlob Frege’s between reference, the object or the “what” purportedly indicated by the question’s sentential subject, and sense, the articulated interrogative sentence itself or at least its logical content.69 We

Hermeneutic Priority, Phenomenological Indeterminacy  235 have restricted our use of the word “meaning” to sense and not reference: judgments only make sense in light of questions. Answers do consist of both sentential subjects (that enact the identifying function of language) and predicates (that enact the elucidatory function), yet identification is not equivalent to reference. In our naïve attitude, a sign purportedly refers to a signified.70 When we ask about “the meaning” of a sentence in this attitude, we ask what object the sentential subject stands for in “external” reality. Reference dominates as the normative definition of meaning or what meaning truly “means.”71 Irigaray criticizes this hegemony as freezing the flow of lived experience, forcing moments into submission to our calculative will.72 Moreover, this makes it seem as if questioning is concerned about extralinguistic reality. We believe our questions are not about “mere” words but about “real” things. Nevertheless, reference is not a meaning in our phenomenological epoche; reference is not an answer to a question.73 Bracketing naïve reference exhibits the dimension of meaning, or sense in its broadest application—like a sense of direction, a sense of ritual, or common sense.74 Sense displaces the supposed immediate character of “sensation” of things to which we believe we “refer” to where we can speak of and interpret experience in response to our dehiscent questionings. To discuss experience is to dignify it by raising it to the light of the logos.75 Genuine questioning makes it possible to receive sense, for judgments to appear as meanings. Meanings are voiced in language, are what we hear spoken to us.76 Not merely sentences perceived, but sentences understood are language. As Gadamer emphasizes, “Being that can be understood is language.”77 Only sentences are understandable. This may seem to atrophy understanding, because in our natural attitude we believe we ought to understand all things, especially “mind-independent” things, and not “merely” language. We express disappointment at not understanding certain things. For example, as Talal Asad writes of the horror of suicide bombing, Breaking into this paranoid [frenzy] may be the sudden realization that in any death there is nothing to understand—that there’s no role for the meaning-making subject. The thought that makes chance deaths more horrible is that they cannot be redeemed by a comforting story.78 Or, others exclaim: “You will never understand my experience!” It seems we want to understand these things, but despair that we cannot. If only sentential answers to asked questions can be understood as meanings, then this despair is not a problem of understanding proper but a problem of our expectations that try to smear understanding across all things. We want them to be meanings when they are not. We cannot understand these things because they are not meaningful judgments. They are neither determinate nor indeterminate in our sense; they do not belong to the order of predicative indeterminacy. We participate in events, undergo experience, and meet others, and these generate a need to speak up.79 We understand discourse—no

236  Nathan Eric Dickman more, no less. These grammatical distinctions serve to protect others from naïve assimilation and to protect understanding from pretentious inflation. Sense is articulate in complete judgments responding to questions. What is the “sense” of genuine questioning? Since the noematic pole of judging is the judgment made, the unity of a subject and a predicate, it may seem that the noematic pole of questioning consists of their separation.80 Let us examine the unity in judgments more closely. The unity in a judgment is not merely the ascription of a property to the fixed base of a thing referenced by a subject. Such a reified judgment, abstracted from the fertilization and irrigation of questioning, is the erosion of living thinking.81 It is merely representational or imagistic, laying one image over another. Instead, the unity in meaningful judgments is where the predicate discloses the subject. Gadamer illustrates this, “ ‘God is one’ does not mean that it is a property of God’s to be one, but that it is God’s nature to be unity.”82 The subject “passes into” and is superseded by the predicate. The judgment does not just state something about something; it presents a conceptual unity by disclosing aspects of the subject’s essence. We can clarify this further through Ricoeur’s semantics of metaphor.83 Metaphors do not just replace a conventional name with an unconventional name—this is representational thinking. Instead, metaphors are sentential, where the literal unity implodes so that the metaphorical unity can emerge. To say, “Paul is a lion,” is to state literal nonsense, but this opens up the metaphorical dimension of meaning whereby the subject’s essence is disclosed. Similarly, to say, “The year is 1438 AH,” discloses this moment as what it truly is, that the essence of time is defined by this moment of ultimate significance when Muhammad founded the Ummah (community) in Medina.84 Sentential subjects set elements of experience into relief in order that we can make a determination about them with a predicate.85 This subtle disclosive relation between subject and predicate is key in judging. What is the nature of this relation in genuine questioning? If a judgment presents an actual unity (or separation) of subjects and predicates, then a question presents not the separation but the possibilities for unity (or separation) of subjects and predicates. As Husserl writes, Every possible content of judgment is thinkable as the content of a question. In the question, it is naturally not yet an actual content; rather, it is in the question only as contemplated, a merely represented (neutralized) judgment and is, as the content of the question, oriented equally toward the yes and the no.86 Genuine questioning presents this indeterminacy, this neutralization of determined judgments. There are a number of orientations of questioning coordinated by different emphases on predicates or subjects.87 Many questions focus on predicative indeterminacy. They contain an articulated subject and open predicative possibilities. For example, “Where are my

Hermeneutic Priority, Phenomenological Indeterminacy  237 car keys?” The sentential subject is explicit (“keys”) but the predicate is yet to be determined. The keys may be in any number of places, and each possible place is a unique predicate. These are predicament-centered questions.88 Other questions focus on indeterminacy of subjects. They contain articulated predicates, but open possible subjects. For example, “What is for dinner?” The predicate is articulated (“is for dinner”), but the subject is yet to be determined. The interrogative pronoun functions similarly to an algebraic variable. It could be “sushi,” “nothing,” and more. These are subject-centered questions. The copula orients a further kind of questioning. These questions presuppose precedent judgments. For example, “Donald J. Trump is the President of the United States. But is he really?” In our natural attitude, we take this as about “mind-independent” reality, because it seems answerable only by looking at the current executive branch of government to see whether Trump is there. In our epoché, questioning is not about “external” reality. What is asked about is the “being” of the relation(s) between the subject and the predicate, namely, the copula. Is the predicate actually disclosive of the subject, or vice versa?89 These are copula-centered questions. Each orientation involves a unique indeterminacy. In predicamentcentered questioning, a subject in relief lacks a definitive determination with a predicate. For example, imagine asking, “What is the distance to Asheville, NC?” The drive to Asheville—the subject—is clear, but a fitting predicate is yet to be determined. The predicates are not vague, because numerous precise measurements are available. The indeterminacy is solely in which predicate applies. In subject-centered questioning, the predicate is in relief, but the subject is not yet determined. For example, “What gets wetter as it dries?” The predicate is clear. Numerous subjects are available as fitting for the predicate—an umbrella, a towel, etc. The indeterminacy is not in the subjects, but in which subject. In copula-centered questioning, a question presupposes an already determined complete thought, so the indeterminacy adheres neither to the subject nor to the predicate poles. For example, “Should one really ‘treat others as one would want to be treated’?” At stake is the justification for the judgment.90 Perhaps it is received convention, divine revelation, or a rational principle. What is indeterminate is which way we will preserve the reliability of our cognitive capacities.91 Whatever orientation, genuine questioning does not say anything about something; judgments do. Nevertheless, it does make something present for us. As Beauvoir emphasizes, intentionality is not a wanting to be, but a wanting to disclose.92 We make the world present to us. Questioning is the way in which we make present the fluidity of subject matters that radiate predicative possibilities and of predicates that radiate possible subject matters. The intentional correlate of questioning is not a “mind-independent” state of affairs, but a conscious aspect of states of affairs present only in questioning. It is, as Johannes Daubert calls it, the Frageverhalt—the state of affairs in the aspect of questionability.93 As exploration of predicate and subject possibilities, questioning gets a train of thought or dialogue going,

238  Nathan Eric Dickman a movement that sometimes might go from a more open-ended and indeterminate Frageverhalt to a more closed and determinate Erkenntnisverhalt (state of affairs as known), but it need not. Questioning is an “in-between,” mediating isolated experience and shared understanding.94 We can share a determination having first shared an indeterminacy. A further orientation of questioning reflects what we are doing here: questioning-centered questioning. At stake in questioning is not only predication of subjects or reliability of pre-judgments. We can also be oriented toward ourselves—not as the subject matter but as participants in questioning. Questioning envelops itself because we disclose something about ourselves, the indeterminacy of “us” examined above. This is what MerleauPonty calls questioning “to the second power.”95 In giving ourselves over to genuine questioning, we disclose ourselves as a negativity “borne by an infrastructure of being.”96 Yet we who question genuinely are not a privative nothing or a “lack.” In giving ourselves to genuine questioning, we let ourselves be disclosed as both distanciated from and inextricably tied to being. We ourselves are suspended in indeterminate yet productive abeyance between distanced negativity and immediate positivity, multiplicity and unity. This radical ambiguity is, in Ricoeur’s words, “the ruin of the pretension of the ego to be established as an ultimate origin.”97 Being in question, we negate our enclosing need for dominance because we are open to what others have to say. Questioning is, in Gadamer’s words, “not merely a matter of putting oneself forward and successfully asserting one’s own point of view, but being transformed into a communion in which we do not remain what we were.”98 Questioning is an opening starting from which it is possible to listen.99 Questioning opens us to possibilities of understanding meanings that may transform us. This openness involves a reworking of experience and interpretations of it, but even more so a reworking of ourselves.100 Such is the reflexive existential indeterminacy in questioning-centered questioning. Let us conclude by highlighting a few areas for further application.

Areas for Further Research Our development of genuine questioning challenges theories that subordinate questioning to propositions. We have bracketed the privative need for answers to expose questioning’s productive surplus of semantic indeterminacy. We have also responded to criticisms that suspect questioning is intrinsically exploitative.101 In genuine questioning, the positionality of questioners is indeterminate and without domination. A further approach to questioning in hermeneutic phenomenology is Gadamer’s characterization of questions as “breaking open” the subject matter.102 This final contrast will help bring our approach into better relief. Gadamer writes, “Discourse that is intended to reveal something requires that that thing be broken open by the question.”103 What does “breaking” something open mean? Questions bring things into a state of indeterminacy

Hermeneutic Priority, Phenomenological Indeterminacy  239 where the answer is not yet settled.104 Gadamer explains, “questioning makes the object and all its possibilities fluid.”105 The aim of questioning is not to figure things out with an answer, but to open up a fluidity of possibilities. I imagine Gadamer’s model here like cracking open an egg, resulting in release of the egg white and yolk from the shell. It opens up a number of possibilities. Does genuine questioning really “break open” the subject matter? Perhaps Gadamer is speaking about doubt instead of genuine questioning. To clarify, consider by comparison how Robert P. Scharlemann, like Husserl, identifies questioning with the essence of doubt.106 Scharlemann defines questioning as a “movement disrupting the object to which it is directed. A question shakes the given structure of the reality with which it is dealing.”107 Doubting-questions dislodge the copula constituting judgments. Yet doubt asserts the “negation” of an affirmative judgment.108 While doubt might break things open, genuine questioning does not. Genuine questioning suspends even negative judgments. It is more fundamental than doubt or assertion. Questioning suggests multiple syntheses of subjects and predicates. In this way, questioning makes a place where positions can occur rather than asserts from a position.109 It embodies an openness to doubt without clinging to negation. This is because with us there are no “shells” that need to be broken open. Our bodily dehiscence is already an openness. We do not deliberately break anything with genuine questioning to make them indeterminate; indeterminacy is disclosed in questions that occur to us. Do we break open others with genuine questions, though? Maybe when we try to force others to “spill their guts.” Interrogation circumvents the other’s autonomy as a genuine conversant, like a witness on the stand. Our focus is genuine questioning, however, and not “breaking” another open. Rather, genuine questioning holds possibilities in suspense, and generates a space in which to welcome what the other says. Genuine questions invite openness rather than impose openness. Perhaps we “break” ourselves open. Through suspension of prejudice, we listen and what we understand may change us. Yet this is not really a breaking or splitting open. Genuine questioning is about renewing who and what we already are in dehiscent openness: homo interrogans.110 There are further areas of application. Three practical interests I focus on are classroom discussions, reading pedagogy, and interreligious dialogue. How might educators help students genuinely question?111 How does shared questioning aid readers’ comprehension of texts?112 How will interreligious understanding change with genuine questioning?113 Beyond these practical applications, one pressing philosophical problem concerns what Rebecca Kukla calls “the transcendental enabler of responsibility.”114 It seems that questioning enables us to determine things for ourselves within horizons of indeterminacy, and thereby take responsibility for our answers. Is not questioning a way we make, a way we determine, indeterminacy—however much we might stress the occurrent character of questioning? How can we take responsibility for our questioning if to take responsibility, to determine

240  Nathan Eric Dickman something as “the” answer, takes antecedent questioning? There must be a responsibility that precedes answering, for surely we are also responsible for our questioning. Some claim that we are “called” into questioning. As Vincent Blok writes, “[W]e have to admit that all questioning presupposes a call: what does not have such a call on us is not questionworthy.”115 Yet this appeal to responsibility is, as Levinas alternatively seems to argue at times, “an interrogation that, behind responsibility and as its ultimate motivation, is a question about the right to be.”116 What is the discursive mood of this transcendental enabler of our ability to respond? Is it a “calling” or a “questioning”?

Notes 1 See Ricoeur (1975). 2 Morgan and Saxton (2006). 3 Merleau-Ponty (1968, pp. 103–107). 4 Raphael (1986). 5 Blok (2015). 6 Plotka (2012). 7 Åqvist (1965). 8 I call these “open-ended” elsewhere (Dickman 2009b, c). 9 Or at least the word “God.” Levinas writes, “Knowledge, the answer, the result . . . would be from a psychism still incapable of thoughts in which the word God takes on meaning.” Levinas (1988a, p. 75). 10 Levinas (1969, p. 102). 11 Like verbs with the same prefix (to describe), to determine is not the removal (unlike to decapitate) but the pinning down of terms. 12 Welton (1999, pp. 366–369). 13 Ricoeur (1975, p. 100), my brackets. 14 Weiss (2008, p. 37). 15 See Schumann and Smith (1987), and Harrah (1961). 16 Husserl (1975, pp. 308–309). 17 Schumann and Smith (1987, p. 372). They continue, “silent interrogations . . . turn out to be the nonstandard or derivative formations.” 18 See Gadamer (1989, p. 47). 19 Based on Gadamer: “To understand a question is to ask. To understand meaning is to understand it as the answer to a question” Gadamer (2013, p. 383). 20 The English “meaning” is polysemic. We look up the “meaning” of words in dictionaries. Lexical entries are definitions. Words do not have a “meaning” because single words and definitions are not complete thoughts or complete answers to questions. We want to know the “meaning” of our lives. This concerns purpose. Life does not have a “meaning” because it does not answer a question. We ask about another’s “meaning” to check intention. An intention is not a “meaning” because it is an existential disposition, not a sentence answering a question. And so forth for reference, significance, etc. We will return to this below. 21 Gadamer (2013, p. 383). 22 See Levinas (1988b). 23 See Dickman (2015). 24 Based on Gadamer’s ontology of play (2013, p. 111). See also Searle (1992, p. 22). 25 Simon (1989). 26 Žižek (1989, pp. 178–182). See also Fiumara (1990, pp. 28–49).

Hermeneutic Priority, Phenomenological Indeterminacy  241 27 Levinas writes, “The other to whom the petition of the question is addressed does not belong to the intelligible sphere to be explored. He stands in proximity” (1998b, p. 25). 28 Irigaray (2002, p. 36). 29 Tillich (1964, pp. 204–205). 30 Booth (1979, pp. 238–239). On complicity with imperialism, see Masuzawa (2005). 31 Merleau-Ponty (1968, p. 103). See Meyer (1995, p. 210). 32 Meyer (1995, p. 126), and Ricoeur (1976, p. 10). 33 See, for example, Bassham and Irwin (2005). 34 Heidegger (1996, p. 166). 35 A subject matter, not a “thing-in-itself.” See Kant’s rejection of “noumena” (2007, pp. 260–261). See Dignaga and Dharmakirti’s approach to “apoha” or exclusion under the condition of shunyata or emptiness in Gillon (2013, p. 315). 36 See Wittgenstein (2009). 37 See Weiss (2008, pp. 58–59). 38 Ricoeur (1975, p. 96). See Sartre (1991). 39 Gadamer (2013, p. 489). 40 See Hegel on the vacuity of “sense-certainty” (1977, p. 58). 41 Ricoeur (1979, p. 129). 42 Ibid., p. 130. 43 Ibid., p. 134. 44 Welton (1999, p. 93). 45 Merleau-Ponty (1968, p. 121). See Plotka (2012). 46 Welton (1999, p. 96); and Sartre (1991, p. 38). 47 Welton (1999, p. 95). 48 Schumann and Smith (1987, p. 370); and Sartre (1991, p. 58). 49 Merleau-Ponty (1968, p. 129). 50 Ibid., p. 121. 51 Merleau-Ponty (1968, pp. 117 and 123). 52 Merleau-Ponty (2002, p. 373). The body is “the visible seer, the audible hearer, the tangible touch—the sensitive sensible: inasmuch as in it is accomplished an [interdependence] of sensibility and sensible thing” (Merleau-Ponty 1968, p. liv). 53 Merleau-Ponty (1968, p. liv). 54 Gadamer (2013, p. 289). 55 Heidegger (1996, pp. 108–122). See Adorno’s criticism (2003, pp. 83–86). 56 Sartre (1991, p. 58). 57 Irigaray (2002, p. 74). 58 Gadamer (2013, p. 310). On the distance in intentionality, see Beauvoir (2015, pp. 10–11). Cf. Bruin (2001, p. 39). 59 Gadamer (1989, p. 26). 60 See Derrida’s examination of language’s opacity undermining exact reproduction of sense against Husserl: Derrida (1973, pp. 121–122). See Adorno’s similar criticism of Heidegger: Adorno (2003, pp. 10 and 47). 61 Meyer (1995, p. 216). Schumann and Smith explain questioning does not objectify; “it does not itself say anything about these objects . . .” (1987, p. 356). 62 Heidegger (1996, p. 3); Searle (1979, p. 14); Harrah (1982, pp. 26–27). 63 Levinas (1988a, p. 71). 64 Ibid., p. 72. 65 Ricoeur calls this “dis-appropriation” (1975, p. 95). Cf. Irigaray (2002, p. 36). 66 Meyer (1995, pp. 204–205). 67 Plato (1991, p. 73 [394d]). 68 See Bell (1975, p. 198).

242  Nathan Eric Dickman 69 Frege (1998). 70 See Derrida’s criticism of Husserl’s notion of “indication” (1973, p. 17). 71 For illustration, “meaning” is used to translate “Bedeutung” (reference) throughout Wittgenstein (2009). 72 Irigaray (2002, p. 40). 73 See Dickman (2016) on the problem of reference and representation. 74 Ricoeur (1975, p. 97). See also Bell (2009, pp. 69–117). 75 Ricoeur writes, “Exteriorization and communicability are one and the same thing for they are nothing other than the elevation of part of our life into the logos of discourse. There the solitude of life is for a moment, anyway, illuminated by the common light of discourse” (1976, p. 19). 76 All meaningful sounds are voices. Idhe writes, “In [a] broad sense one may speak of the voices of significant sound as the ‘voices of language.’ ” Idhe (2007, p. 147). 77 Gadamer (2013, p. 490). 78 Asad (2007/2008, p. 129). 79 See Ricoeur (1975, p. 99). We need not be worried about “linguistic solipsism” or “linguistic idealism.” See Lawn (2006), 84. Questioning is how we, Merleau-Ponty remarks, “transform the powers of death into poetic productivity” (1968, p. 116). 80 This is not so. See Schumann and Smith (1987, p. 365). 81 See Gadamer on Hegel’s approach to the proposition: Gadamer (2013, p. 482). 82 Ibid. 83 Ricoeur (2003). 84 To say, “The year is 2017 CE” discloses time’s essence with regard to the birth of Jesus. Further still, Christians believe that the essence of Jesus is disclosed by the predicate “Christ.” See Dickman (2017). 85 Schumann and Smith (1987, p. 367). Any element of the lifeworld may be set into relief. To say anything can is not to say “everything” can. In Wittgensteinian fashion, “everything” does do what subjects do. Kant develops the inverse of this in explaining that “existence” cannot function as a predicate. Kant (2007, p. 504). 86 Husserl (1975, p. 309), emphasis added. 87 Bruin distinguishes between “predicative questions” and “hermeneutical questions.” Bruin (2001, pp. 20–24). 88 Because they concern the predicament that the subject is in. This is drawn from Kant’s description of the fundamental categories as “predicaments.” Kant (2007, p. 106). 89 Husserl (1975, p. 294). 90 Ibid. 91 For further development of this sort of questioning, see Husserl (1975, pp. 294–313). 92 Beauvoir (2015, p. 11). 93 Schumann and Smith (1987, p. 369). 94 Questioning marks an “in-between” of predicative fixation and the continual self-correction of consciousness continually taking place in pre-predicative experience. See Weiss (2008, p. 137); and Husserl (1975, p. 294). 95 Merleau-Ponty (1968, p. 120). 96 Ibid. 97 Ricoeur (1975, p. 95), emphasis added. 98 Gadamer (2013, p. 387). 99 Irigaray (2002, p. 36). 100 Paraphrased from Beatty (1999, p. 295). 101 See Comay (1991, p. 149).

Hermeneutic Priority, Phenomenological Indeterminacy  243 102 See Gadamer (2013, p. 383); and Dickman (2009c, pp. 160–162). See also Heidegger’s formulation of questioning (1996, p. 4). I address this in Dickman (2009c, pp. 104–109); and Dickman (2009a). 103 Gadamer (2013, p. 383). 104 The thing, writes Gadamer, “has to be brought into this state of indeterminacy, so that there is an equilibrium between pro and contra” (2013, p. 383). 105 Ibid. 106 Scharlemann (1969, p. 137); and Husserl (1975, pp. 294–311). 107 Scharlemann (1969, p. 137). Breaking something open separates even the fulfilling intuition from the intentional conception, making both the predicative possibilities and the very structure of the subject matter itself fluid. 108 See Merleau-Ponty (1968, p. 120), for a critique of both doubt and positivism as intrinsically nihilistic. 109 Silverman (1986, p. 88). 110 See Bruin (2001). 111 I address this in Dickman (2009b). 112 See Raphael (1986). I address this in Dickman (2014). 113 I address this in Dickman (2015). 114 Kukla (2002). 115 Blok (2015, p. 321). Elsewhere I have challenged the notion of “call” that others, such as Levinas and Heidegger, emphasize as the transcendental enabler of responsibility. See Dickman (2017, online). 116 Levinas (1998a, p. 169).

Works Cited Adorno, Theodor. 2003. The Jargon of Authenticity, translated by K. Tarnowski and F. Will. New York, NY: Routledge. Åqvist, Lennart. 1965. A New Approach to the Logical Theory of Interrogatives. Uppsala: University of Uppsala. Asad, Talal. 2007/2008. “On Suicide Bombing.” The Arab Studies Journal 15(2)/16(1): 123–130. Bassham, Gregory. and W. Irwin. 2005. Critical Thinking: A Student’s Introduction, 2nd edition. New York, NY: McGraw Hill. Beatty, Joseph. 1999. “Good Listening.” Educational Theory 49(3): 281–298. Beauvoir, Simone. 2015. The Ethics of Ambiguity, translated by Bernard Frechtman. New York, NY: Open Road. Bell, Catherine. 2009. Ritual Theory, Ritual Practice. Oxford: Oxford University Press. Bell, Martin. 1975. “Questioning.” The Philosophical Quarterly 25(100): 193–212. Blok, Vincent. 2015. “Heidegger and Derrida on the Nature of Questioning: Towards the Rehabilitation of Questioning in Contemporary Philosophy.” Journal of the British Society for Phenomenology 46(4): 307–322. Booth, Wayen C. 1979. Critical Understanding: The Powers and Limits of Pluralism. Chicago: The University of Chicago Press. Bruin, John. 2001. Homo Interrogans: Questioning and the Intentional Structure of Cognition. Ottawa, ON: The University of Ottawa Press. Comay, Rebecca. 1991. “Questioning the Question: a Response to Charles Scott.” Research in Phenomenology 21(1): 149–158. Derrida, Jacques. 1973. Speech and Phenomena, translated by D. Allison. Evanston, IL: Northwestern University Press.

244  Nathan Eric Dickman Dickman, Nathan Eric. 2009a. “Anxiety and Face of the Other: Tillich and Levinas on the Origin of Questioning.” Sophia: International Journal for Philosophy and Traditions 48(3): 267–279. ———. 2014. “Between Gadamer and Ricoeur: Preserving Dialogue in the Hermeneutical Arc for the Sake of a God Who Speaks and Listens.” Sophia: International Journal for Philosophy and Traditions 53(4): 553–573. ———. 2017. online. “Call or Question: A Rehabilitation of Conscience as Dialogical.” Sophia: International Journal for Philosophy and Traditions. ———. 2009b. “The Challenge of Asking Engaging Questions.” Currents in Teaching and Learning. 2(1): 3–16. ———. 2009c. Dialogue and Divinity: A Hermeneutics of the Interrogative Mood in Religious Language. Ph.D. diss., The University of Iowa. ———. 2015. “Ethical Understanding: The Priority of Questioning in Interreligious Dialogue.” Listening: Journal of Communication Ethics, Religion, and Culture. 50:2 (Spring): 92–105. ———. 2016. “Linguistically Mediated Liberation: Freedom and Limits of Understanding in Thich Nhat Hanh and Hans-Georg Gadamer.” The Humanistic Psychologist. 44: 256–279. ———. 2017. “Transcendence Un-Extra-Ordinaire: Bringing the Atheistic I Down to Earth.” Religions. 8(4): 1–21. Fiumara, Gemma Corradi. 1990. The Other Side of Language: A Philosophy of Listening, translated by C. Lambert. New York, NY: Routledge. Frege, Gottlob. 1998. “On Sense and Meaning,” in Philosophy of Language: The Big Questions, edited by A. Nye (pp. 72–76). Malden, MA: Blackwell. Gadamer, H. 1989. “Text and Interpretation,” in Dialogue and Deconstruction: The Gadamer-Derrida Encounter, edited by D. Michelfelder and R. Palmer (pp. 21–51). Albany, NY: SUNY Press. ———. 2013. Truth and Method, 2nd revised edition, translated by J. Weinsheimer and D. Marshall. London: Bloomsbury Academic, Reprint Edition. Gillon, Brendan S. 2013. “Language and Logic in Indian Buddhist Thought,” in A Companion to Buddhist Philosophy, edited by S. Emmanuel (pp. 307–319). Chichester: Wiley-Blackwell. Harrah, David. 1961. “A Logic of Questions and Answers.” Philosophy of Science 28(1): 40–46. ——. 1982. “What Should We Teach about Questions?” Synthese. 51: 26–27. Heidegger, Martin. 1996. Being and Time: A Translation of Sein und Zeit, translated by J. Stambaugh. New York, NY: SUNY Press. Hegel, G. W. F. 1977. Phenomenology of Spirit, translated by A. Miller. Oxford: Oxford University Press. Husserl, Edmund. 1975. Experience and Judgment, translated by J. Churchill and K. Ameriks. Evanston, IL: Northwestern University Press. Idhe, Don. 2007. Listening and Voice: Phenomenologies of Sound, 2nd edition. Albany, NY: SUNY Press. Irigaray, Luce. 2002. The Way of Love, translated by H. Bostic and S. Pluhacek. New York, NY: Continuum. Kant, Immanuel. 2007 [Original 1781]. Critique of Pure Reason, translated by M. Weigelt and M. Muller. London: Penguin Classics. Kukla, Rebecca. 2002. “The Ontology and Temporality of Conscience.” Continental Philosophy Review. 35: 1–34.

Hermeneutic Priority, Phenomenological Indeterminacy  245 Lawn, Chris. 2006. Gadamer: A Guide for the Perplexed. New York, NY: Continuum. Levinas, E. 1988a. “Hermeneutics and Beyond,” in Entre Nous: On Thinking-of-theOther, translated by M. Smith and B. Harshav (pp. 65–76). New York, NY: Columbia University Press. ———. 1988b. “Nonintentional Consciousness,” in Entre Nous: On Thinking-of-theOther, translated by M. Smith and B. Harshav (pp. 123–132). New York, NY: Columbia University Press. ———. 1998a. “Notes on Meaning,” in Of God Who Comes to Mind, translated by B. Bergo (pp. 152–171). Stanford: Stanford University Press. ———. 1998b. Otherwise than Being or Beyond Essence, translated by A. Lingis. Pittsburgh, PA: Duquesne University Press. ———. 1969. Totality and Infinity: An Essay on Exteriority, translated by A. Lingis. Pittsburgh, PA: Duquesne University Press. Masuzawa, Tomoko. 2005. The Invention of World Religions: Or, How European Universalism was Preserved in the Language of Pluralism. Chicago: The University of Chicago Press. Merleau-Ponty, M. 2002. Phenomenology of Perception, translated by C. Smith. New York, NY: Routledge. ———. 1968. The Visible and the Invisible, translated by A. Lingis. Evanston, IL: Northwestern University Press. Morgan, Norah and Juliana Saxton. 2006. Asking Better Questions, 2nd edition. Markham, ON: Pembroke. Meyer, Michel. 1995. Of Problematology: Philosophy, Science, and Language, translated by D. Jamison. Chicago: University of Chicago Press. Plato. 1991. The Republic, 2nd ed., translated by Alan Bloom. New York, NY: Basic Books. Plotka, Witold. 2012. “Husserlian Phenomenology as Questioning: An Essay on the Transcendental Theory of the Question.” Studia Phenomenologica. 12: 311–329. Raphael, Taffy E. 1986. “Teaching Question Answer Relationships, Revisited.” The Reading Teacher 39(6): 516–522. Ricoeur, Paul. 1979. “The Function of Fiction in Shaping Reality.” Man and World 12(2): 123–141. ———. 1976. Interpretation Theory: Discourse and the Surplus of Meaning. Fort Worth, TX: Texas Christian University Press. ———. 1975. “Phenomenology and Hermeneutics.” Nous 9(1): 85–102. ———. 2003. The Rule of Metaphor, translated by R. Czerny. London: Routledge Classics. Sartre, Jean-Paul. 1991. The Transcendence of the Ego: An Existentialist Theory of Consciousness, translated by F. Williams and R. Kirkpatrick. Hill and Wang. Scharlemann, Robert P. 1969. Reflection and Doubt in the Thought of Paul Tillich. New Haven, CT: Yale University Press. Schumann, Karl and Barry Smith. 1987. “Questions: An Essay in Daubertian Phenomenology.” Philosophy and Phenomenological Research 47(3): 353–384. Searle, John R. 1992. “Conversation,” in (On) Searle On Conversation, edited by H. Parret and J. Verschueren (pp. 7–30). Philadelphia, PA: John Benjamins. ———. 1979. Expression and Meaning: Studies in the Theory of Speech Acts. Cambridge: Cambridge University Press. Silverman, H. 1986. “Hermeneutics and Interrogation.” Research in Phenomenology 16(1): 87–94.

246  Nathan Eric Dickman Simon, Josef. 1989. “The Good Will to Understand and the Will to Power,” in Dialogue and Deconstruction: The Gadamer-Derrida Encounter, edited by D. Michelfelder and R. Palmer (pp. 162–175). Albany, NY: SUNY. Tillich, Paul. 1964. “Communicating the Christian Message: A Question to Christian Ministers and Teachers,” in Theology of Culture (pp. 201–215). Oxford: Oxford University Press. Weiss, Gail. 2008. Refiguring the Ordinary. Bloomington: Indiana University Press. Welton, Donn. 1999. The Essential Husserl: Basic Writings in Transcendental Phenomenology. Bloomington: Indiana University Press. Wittgenstein, Ludwig. 2009. Philosophical Investigations, 4th revised edition, translated by G. Anscombe, P. Hacker, and J. Schulte. Oxford: Blackwell Publishing. Žižek, Slavoj. 1989. The Sublime Object of Ideology. New York, NY: Verso.

12 Against the Darkness Beauty and Indeterminacy in John Williams’s Stoner Phillip E. Mitchell

John Williams’s novel Stoner is a focused literary study in restraint and understatement, which, in all probability, is the reason for its longstanding obscurity. Despite the number of authors and book reviewers who have praised Williams’s work for its depiction of academic life, its beautiful but painful stoicism, and its portrayal of a life in letters, its first run sold only two thousand copies, and it has spent several decades going in and out of print.2 In recent years, however, the tides have begun to change, and the novel has begun experiencing a “surprising revival,” selling thousands of copies in the US, throughout Europe, and as far away as Israel.3 That it has seen renewed popular interest in recent years is, indeed, a surprise, not least of all because its subject matter is unlikely for a hit novel: a college professor’s life, from beginning to end, with unrelenting tragedy between. William Stoner is a victim, perhaps perpetrator too, of a passionless marriage; his wife is obstinate and cold; he offends a colleague early in his career and pays for it the rest of his life; he has one great love affair with a student who is forced to leave the university at the hands of the vindictive colleague; and, finally, he dies an anonymous and unappreciated figure. In The New Yorker, Tim Kreider writes, correctly, that Stoner’s life is “a bummer.”4 But the experience of reading Stoner is not a bummer at all. C. P. Snow writes, “It sounds grey but it isn’t in the least grey, because of the spirit behind it . . . The curious thing is, and it is a triumph both for Mr. Williams’s humanity and his art, that one ends the book with a totally different feeling.”5 Snow suggests the book “produces a kind of pride.”6 Morris Dickstein, likewise, believes “it is a perfect novel, so well told and beautifully written, so deeply moving, that it takes your breath away.”7 And, in a glowing piece about the novel, Mel Livatino writes, “I hope I haven’t sounded too unhinged in my appreciation, but that is what this novel can do to its reader.”8 Despite gaining a wide international readership over the past decade, and spawning myriad online reviews, few scholars have written at length about this odd dynamic between subject matter and the experience of the novel.9 Tony McKenna is the only critic to offer extended critical commentary on the form of the novel.10 He suggests that the novel’s power “lies in its

248  Phillip E. Mitchell articulation—with bare, beautiful artistry—of the most fundamental contradiction that nestles at the heart of being: the aporia between the transitory and the infinite, the mundane and the ethereal—the vertiginous sense of the individual stood on the edge of forever.”11 To McKenna, the novel is “anything but” nihilistic because of the transformative power of love that comes through Stoner’s lifelong study of literature.12 The transcendent, or infinite, is mediated through Stoner’s love of his vocation.13 McKenna, however, sees the form and substance of the novel as one and the same. The “bare, beautiful artistry” of the novel is akin to the bleakness of its subject matter.14 I, on the other hand, see the style of the narrative and the subject matter as distinct elements in the text.15 Each might seem on the surface to negate the other, but I want to show how the two impulses, in their opposite positions, come together by the end of the work to form the unity of the text. This unity, however, is not unity in the sense that indeterminacy between form and content has been dispelled; rather, the two elements work against each other, propelling and completing the narrative. In this way, the novel achieves what the New Critics call “organic unity” or what John Keats calls “negative capability.” The total vision of the poet is one that can hold oppositions together in a single vision. For our purposes in this collection, the phrase “indeterminate unity” might be a proper substitute. Further, I want to show the extent to which the work is a comment on these very processes, becoming a parable about the role of reading in meaningmaking within an indeterminate reality. I believe, unlike McKenna, that the novel is self-referential, suggesting that what elevates one’s experience, or makes one’s experience of life worth having, is, according to the novel, not the experience of the transcendent, but, rather, the experience of art itself.16 In this way, the novel points to the importance of art to represent the “experience of art” as a hedge against existential despair. I agree with McKenna that it is love that transforms the narrative, and Stoner, but it is a love that circles back on itself and offers a commentary on the very nature of “reading” and the read work. The interplay between the poetic impulse and the despairing subject matter works to effect this phenomenon. The reader must, just as Stoner does, learn to hold the opposing forces together. By embracing the indeterminacy inherent in the differences between a world without apparent meaning and the poetic exploration of that reality, the reader, like Stoner, is able to forge a new kind of self, one formed by the aesthetic experience. A close look at the trajectory of William Stoner’s life reveals the nihilistic subject matter of the novel. The titular character lives up to his namesake, analogous to a statue—expertly crafted, appearing to move within the world while simultaneously stationary in it and separated from it by his essence. Even the name of the novel points to his lack of volition, or his inability, to “act” in the world. And, it is not as if it is because he is Fyodor Dostoevsky’s overly conscious subject. In fact, we know little

Against the Darkness  249 about the inward life of William Stoner. The few clues that we have regarding his consciousness come mediated through the third-person narration. What we do know, however, is that he moves specter-like through his life, pushed about by brute forces. In this way, his despair elevates him to the archetypal. There is a story arc but not a character arc, per se, and we see a representation not of individuality at work against circumstances, but, rather, a statue made from, and vulnerable to, the chaos and immutable despair of a world. When he delights in his daughter’s alcoholism, when he responds passively to his wife’s haranguing, when his supervisor (vindictively) offers him only composition classes, he is as remote and removed as his namesake. What is more, the novel begins and ends with his death. In this way, the narrative undermines the typical literary arc. He does not change; he has no great insight that affects his relationships with other people; there is no resolution celebrating a renewed vision of the outside world. It is despair all the way down. Livatino rightfully calls it “a sadness unto the bone.”17 When explaining his role in the memory of his former students and the faculty at the University of Missouri, the lines are terse, matter of fact, and bleak. Stoner’s name is only a “sound” to the younger generation, a reminder of only “the end that awaits them all.”18 The reader confronts the reality head-on that the protagonist with whom he or she will spend the next 278 pages is not heroic in the classical sense. He is no Achilles or Odysseus. No one sings songs about Stoner. His dissertation, his single contribution to academia, lies buried in the library. Given the bleak end that awaits Stoner, the narrator immediately drives the reader into ambiguous territory, wherein he or she is uncertain whether, given Stoner’s anonymity, his life is worth reading. He cannot evoke “the past” in any romantic sense. He is only a statue, present but lifeless. Because the reader holds in his or her hands a work bearing Stoner’s name, however, he or she must believe that, despite his anonymity and his mediocre life, there must be something worth reading about him. Kreider believes such a beginning “audacious” and suggests that “by preëmpting the usual suspense of narrative, denying us even the promise of some cathartic tragedy, Williams forces us to wonder: What will this book be about?”19 Here the novel introduces to the reader the indeterminate space that will inform the rest of the story, inviting the reader to take the challenge of undergoing a journey into finding significance where the endgame has suggested he or she cannot. It cautions the reader: this book might very well be The Inferno without Beatrice. The beginning evokes the fear and dread of living a futile life while enticing the reader with a sense that the endeavor of reading will yield greater returns. Despite the hard surfaces of reality, the novel compels the reader to find meaning in the quotidian. Working against the despair, however, is the movement of the prosody and the technical acumen of the novel. As Snow believes, “This is, in the very best sense, a poet’s novel.”20 A close look at the poetic presences of the

250  Phillip E. Mitchell narrative reveal their work in building an aesthetic at odds with the subject matter. When Stoner visits home after college, He turned on the bare, treeless little plot that held others like his mother and father and looked across the flatland in the direction of the farm where he had been born, where his mother and father had spent their years. He thought of the cost exacted, year after year, by the soil; and it remained as it had been—a little more barren, perhaps, a little more frugal of increase. Nothing had changed. Their lives had been expended in cheerless labor, their wills broken, their intelligences numbed. Now they were in the earth to which they had given their lives; and slowly, year by year, the earth would take them. Slowly the damp and rot would infest the pine boxes which held their bodies, and slowly it would touch their flesh, and finally it would consume the last vestiges of their substances. And they would become a meaningless part of that stubborn earth to which they had long ago given themselves.21 While the passage is bleak, it is also charged with poetic energy. Stoner thinks abstractly of the “cost exacted.” He also thinks of their “Cheerless labor,” “broken wills,” and “numbed intelligences,” indicating a poet’s sensibility in conceptualizing his parents’ lives. Furthermore, the earth is given creaturely properties, an active figure that “infests,” “takes,” and “consumes.” Their bodies are not simply corpses, but, rather, “vestiges of their substances.” Likewise, the earth is “stubborn.” Death gives way to the linguistic imagining of decay. While Stoner’s birthplace is “a little more barren,” the passage itself is rich and complex. Through Shakespeare’s seventy-third sonnet, Stoner himself confronts this very dichotomy between the bleakness of life and the beauty of literature. The meeting quickens stoner’s pulse: William Stoner realized that for several moments he had been holding his breath. He expelled it gently, minutely aware of his clothing moving upon his body as his breath went out of his lungs. He looked away from Sloane about the room. Light slanted from the windows and settled upon the faces of his fellow students, so that the illumination seemed to come from within them and go out against a dimness.22 The final lines of the poem, “This thou perceives, which makes thy love more strong,/To love that well which thou must leave ere long,” suggest the power of love in the face of death. When his teacher, Arthur Sloane, presses Stoner for a response, asking Stoner “What does the sonnet mean?” Stoner is befuddled and unable to answer.23 After the awkward exchange, Stoner abandons his studies in agriculture and devotes himself solely to the study of literature. Material reality, that of life working with the soil in this case, holds no power pitted against the aesthetic for Stoner. He becomes

Against the Darkness  251 “conscious of himself in a way that he [has] not been before.”24 He begins reading voraciously, “inhaling the musty odor of leather, cloth, and drying page as if it were an exotic incense.”25 Literary language becomes a mirror for Stoner, showing himself to himself or, rather, creating a “self” that heretofore has not existed. Although the narrative provides scant exposition for why Stoner reacts in such a way, it is obvious that the reason for his reaction is two-fold: first, he faces death in the poem. Outlined in the subject of the poem are the boundaries of life and the finite nature of experience. Second, he faces life in the sonnet through artistic representations. And, it is in the poem’s representation of dying, its poetic language, that experience is elevated to the level of “meaning.” This, however, is not something he will understand until later. Early on, he sees only “the potentialities of prose and its beauties.”26 Nowhere clearer is this dynamic between poetic language and despair at work than in the interplay of darkness and light. While Williams seldom gives detailed descriptions of people or spatial realities, he often sets scenes by describing the way light falls and its relationship to shadows and the darkness around. It works in two ways throughout the novel, as analogy and as exposition. Ultimately, it is only through language that Stoner finds “light.” For instance, we see the prior approach when his class studies Shakespeare’s sonnet and “illumination [seems] to come from within [the students] and go out against a dimness.”27 And, when he begins his literary studies, as he reads his “lamplight [flickers] against the shadows.”28 Williams writes, If he stared long and intently, the darkness gathered into a light, which took the insubstantial shape of what he had been reading. And he would feel that he was out of time, as he had felt that day in class when Archer Sloane had spoken to him. The past gathered out of the darkness where it stayed, and the dead raised themselves to live before him; and the past and the dead flowed into the present among the alive, so that he had for an intense instance a vision of denseness into which he was compacted and from which he could not escape, and had no wish to escape.29 The poetic passage is dizzying in its sentiment, both uncanny and exhilarating. The past is figured here as the dead rising into light. They surround him, and they leave him suffocated. At the end of the passage, we see that he has no wish to escape the figures. He looks headlong at them. This is the beginning of his studies, his first experience of the terror of existence rising into the beauty of aesthetic experience. Nietzsche, in the first half of The Birth of Tragedy, proposes that Apollo and the Greek god pantheon arose out of a need to respond to Silenus’s proclamation that “What is best [is] not to be born, not to be, to be nothing” or “to quickly die.”30 He believes that “out of the original thearchy of terror the Olympian thearchy of joy gradually evolved through the Apollonian impulse towards beauty, just as roses bud

252  Phillip E. Mitchell from thorny bushes.”31 Here, figured as light and darkness, Stoner experiences the darkness of death and meaninglessness rising into the light of life. There are other examples: Two white candles in dull brass holders glowed unevenly against the darkness”;32 “He left the door to the unlighted bedroom open; the candlelight glowed feebly in the darkness”;33 “iridescent crystals that winked against the grayness”;34 “The image was dark, and he could not make out its features; it was as if he saw a ghost glimmering unsubstantially out of a hardness, coming to meet him”;35 “he sometimes felt that he hunched his back futilely against the driving storm and cupped his hands uselessly around the dim flicker of his last poor match.”36 What is interesting in this relationship is not that the light easily overtakes the darkness. Rather, the light is always threatened by the darkness. What light we do see is winking or dimly glowing throughout the novel, almost always on the verge of petering out. More than his conflicts with Edith (his wife), Lomax (his supervisor), or Charles Walker (his student), this dynamic is the primary conflict of the story and a mirror of the thematic dimension of the work. The mood itself is indeterminate, the light always on the verge of being snuffed out by the darkness but shining nonetheless. It is appropriate, then, that the physical place where Stoner finds light, the university, is set apart from the rest of the world—though not unaffected by it. Discussion of this separation emerges early on in a conversation between Stoner and his friends, Masters and Finch. Masters believes that the university “is an asylum [. . .] a rest home, for the infirm, the aged, the discontent, and the otherwise incompetent”—for the “dispossessed of the world,” people who “would always be on the fringe of success” but “destroyed by [their] failure.”37 Despite this, Masters suggests Stoner has grand delusions, calling him “the dreamer, the madman in a madder world, [their] own Don Quixote without his Sancho, gamboling under the blue sky.”38 To Masters, the “outside” world is a place where their lives devolve into meaninglessness and despair. And, although to Stoner the bleak idea brings “him no vision of the University to which he [has] committed himself,” it seems that for the narrator the “other world” always threatens to upend the safety of the university, just as light and darkness wage a battle for his soul. This outside darkness infiltrates the academy during both World War I and War World II. During WWI, Archer Sloane looks “ten years older,” and his “black eyes had gone dull, as if filmed over with layers of moisture.”39 He frightens students with his eccentricities and stops lectures randomly, refusing to look at his students and remaining silent for up to five minutes at a time.40 Against this backdrop, even Stoner struggles to understand his place in the world. When he considers joining the effort against the

Against the Darkness  253 Germans, he finds “the task of searching his motives a difficult and slightly distasteful one” and feels “that he [has] little to offer to himself and there [is] little within him which he could find.”41 The Great War hovers over the protagonist and his friends, and, maybe even more, Archer Sloane, contributing not only to an uneasiness with the world but also disrupting his sense of inner stability. However, when Masters asks Stoner how he has reached his decision not to enlist, Stoner also, conversely, imagines he has been “miraculously revived” from “the deadness” of the years on his parents’ farm.42 So, which is it? Can Stoner consider himself miraculously revived if there is nothing within him? The answer, of course, lies in the reality that Stoner’s sense of self is determined foremost by aesthetic experience. He is, appropriately, then, exactly the kind of person Masters claims he is; he is Don Quixote, who has started to form his reality based on the books that he has read.43 When he tries to look inside himself in regards to the outside world, the effort dumbfounds him. Nothing exists in Stoner that is not somehow tied to his studies. Livatino believes, “William Stoner can exist only in a university that still values such contemplation [when the university was a retreat from the world]” and that “his teaching excels not because he is brilliant, creative, or flashy . . . but because he is witness to such a consciousness.”44 It is no surprise, then, that he cannot imagine himself outside of that world and decides, like Archer Sloane, that “a scholar should not be asked to destroy what he has aimed his life to build.”45 Stoner’s story is not so much about the world, therefore, as it is the light of aesthetic experience that arises out of that world. His reaction to Masters’ death during the war is telling: When he had thought of death before, he had thought of it either as a literary event or as the slow, quiet attrition of time against imperfect flesh. He had not thought of it as the explosion of violence upon a battlefield, as the gush of blood from a ruptured throat. He wondered at the difference between the two kinds of dying, and what the difference meant.46 Coping with the external realities perplexes Stoner; consequently, he chooses to return to conceptualizing Masters’s death as a literary event, calling him “Catullus or a more gentle and lyrical Juvenal, an exile in his own country.”47 Indeterminacy, then, ad initium, is the necessary condition for the aesthetic experience. The novel makes clear that Stoner’s identity is fluid, perhaps even unformed. One’s response to such a condition has two results: possession by either aesthetic experience or despair. Stoner, like Don Quixote, is the consciousness possessed by the aesthetic, who uses the written word as a basis for the creation of selfhood. Unlike Stoner, his daughter, ironically named “Grace,” does not revel in language and literature. Her

254  Phillip E. Mitchell only means of coping with her failed marriage is alcohol. Thus, her darkness is never mitigated. Instead, she always teeters on the edge of despair: “Grace had become almost motionless, as if she felt that any movement might throw her into an abyss from which she would not be able to clamber”;48 “It was as if something inside her had gone loose and soft and hopeless, as if at last a shapelessness within her had struggled and burst loose and now persuaded her flesh to specify that dark and secret existence”;49 “And Stoner came to realize that she was, as she had said, almost happy with her despair; she would live her days out quietly, drinking a little more, year by year, numbing herself against the nothingness her life had become. He was glad she had that, at least; he was grateful that she could drink.”50 This last quotation is one of the most difficult to read in the novel, for Grace has no escape. She will never bound back, never truly escape from her desolation. Without art, the individual is left with only the hard surfaces of reality. Narrative imposes form upon the chaos in order to make sense of or explore these realities. Through literature, the book tells us, the terror of “nothing” is controlled and its power undermined. The self, always indeterminate— unstable and fluid—meets in the aesthetic experience a mirror that, at least for the duration of the narrative, possesses it, substituting the unstable self for a vision of a “storied” self. It is the experience that Stoner must return to throughout his life in order to survive. This is a literary work, then, about experiencing literary works, producing “within itself” what Linda Hutcheon calls “a dramatized mirror of its own narrative or linguistic principles.”51 While this is not overt metafiction, like the works of John Barth, Kurt Vonnegut et al., because its focus is not only on reading but also making a life out of, or through, what one has read, it is hard to ignore the ways that the novel is a mirror of the very processes that enable it to operate. It speaks to the very nature of the read reality. As a result, the reader and Stoner take a parallel journey to understanding the role of art in life. Because, like Stoner, the reader confronts the bleak determinism of the novel through poetic iteration, he or she is also a student, not of Archer Sloane but, rather, of the narrator, and Archer Sloane’s question, “What does it mean?” is akin to the narrator’s, inviting the reader to conceptualize meaning when it appears there is none. The narrative asks the reader, “What does Stoner’s life mean?” When Sloane reads Shakespeare’s seventy-third sonnet aloud, it is as if he becomes the sonnet, suggesting, as Tony McKenna writes, “you do not possess aesthetic beauty—rather, it possesses you.”52 The movement toward possession of this kind is central to understanding the novel. For Sloane, a seasoned literature professor, poetic language and selfhood meet easily in the sonnet. He, quite frighteningly, finds himself in the grip of the work. He achieves a kind of negative capability, holding the opposites of content

Against the Darkness  255 (the material reality of death) and form (its poetic imagining) in the sonnet together as he reads. According to the novel, this reality is not easily translatable, for it is only the result of a constitution sensitive to the aesthetic experience that can achieve it. Stoner will not experience it fully until the beginning of his affair with Katherine Driscoll, when he learns “that love is not an end but a process through which one person attempts to know another.”53 Through the process of reading, one comes to see, and know, the light of another’s consciousness and, thus, can become “possessed” by the aesthetic experience. It is here that the novel thematizes “the overwhelming power and potency of words, their ability to create a world more real than the empirical one of our experience.”54 Kreider puts it this way: “For all the jewel-like beauty of its own prose, Stoner tells us that the words themselves are inessential; literature, like Stoner himself, is only an imperfect reflector of that light that comes from outside.”55 Midway through his career Stoner sees this overwhelming power of language to give his life meaning, finally displaying “boldly” a “love of literature, of language, of the mystery of mind and heart showing themselves in the minute, strange, and unexpected combinations of letters and words, in the blackest and coldest print.”56 A few lines later, the narrator communicates that Stoner feels he is “at last beginning to be a teacher, which was simply a man to whom his book is true, to whom is given a dignity of art that has little to do with his foolishness or weakness or inadequacy as a man.”57 “Art” is something different from the experience of the everyday, the extratextual or extra-linguistic—reality in its hard shapes. Furthermore, Stoner’s new knowledge is a “knowledge of which he [can] not speak, but one which changed, once he had it, so that no one could mistake its presence.”58 The unutterable and mysterious dimensions of his discipline ground him, once again demarcated from the material or quantifiable. What animates the “blackest and coldest print” is the ineffable power of consciousness as it experiences literature. In the blackest and coldest print of the novel, which encompasses Stoner’s bleak life, the reader, too, elevates Stoner’s life, possessed by a knowledge he or she cannot articulate, other than to say the “book is true.”59 This Romantic concept, perhaps, is the most puzzling indeterminate place in the story. While the novel points to “light,” “beauty,” and “knowing another’s consciousness,” these concepts are impossible to explain fully. The elevated language can point to the transcendent but cannot communicate it definitively. In this way, the novel asks the reader to have a kind of “faith,” both in the authenticity of Stoner’s experience and in the power of literature to engender this kind of response. By emphasizing Stoner’s intimate relationship with language, the novel also self-referentially refers to its own dependence on the reader, who is a mirror of Stoner with the text in his or her hands. Hutcheon writes, “As the novelist actualizes the world of his imagination through words, so the reader— from those same words—manufactures in reverse a literary universe that is as much his creation as it is the novelist’s.”60 There is, then, a space wherein

256  Phillip E. Mitchell the novel cannot signify, cannot “mean,” without the receptor of the representational experience. The narrative without a reader, then, is like Stoner without literature, and the act of reading Stoner is the act of living Stoner’s life, elevating a static and passive text to a lived experience. Hutcheon speaks of this relationship between reader and narrative quite succinctly: [The reader] cannot avoid this call to action for he is caught in that paradoxical position of being forced by the text to acknowledge the fictionality of the world he too is creating, yet his very participation involves him intellectually, creatively, and perhaps even affectively in a human act that is very real, that is, in fact, a kind of metaphor of his daily efforts to “make sense” of experience.61 What enables Stoner to make sense of the world is literature. And, by animating the novel, the reader, too, makes sense of the world (both the world of Stoner and the world of experience) in the same way. Stoner is a great novel, then, not only because of its pitch-perfect descriptions or concise and moving expositions, or its expertly parsed time consolidations, but also because it reveals and paints truthfully the picture of the relationship between the receiver of a literary work and the work itself. It offers technical adeptness and a vision of a particular reality while its autonomic maneuvers circle back upon themselves. In short, the reader understands Stoner because Stoner is a reader. The aesthetic experience that Stoner has is, likewise, the aesthetic experience of the reader. In a piece bemoaning the resurgence of Stoner in literary culture, Drew Smith mockingly, but accurately, describes this phenomenon: “This case has been surprisingly tenacious, I think, partly because Stoner is a book about a literary scholar with a special affection for the written word, which naturally leads to the question, Are all these people who are ‘discovering’ Stoner actually just falling in love with an image they have of themselves?”62 The answer is, of course, “yes.” When reading Katherine Driscoll’s manuscript toward the end of his life, the fragments of Stoner’s narrative piece themselves together into a total vision: Beneath the numbness, the indifference, the removal, it was there, intense and steady; it had always been there. In his youth he had given it freely, without thought; he had given it to the knowledge that had been revealed to him—how many years ago?—by Arthur Sloane; he had given it to Edith, in those first blind foolish days of his courtship and marriage; and he had given it to Katherine, as if it had never been given before. He had, in his odd ways, given it to every moment of his life, and had perhaps given it most fully when he was unaware of his giving. It was a passion neither of the mind nor of the flesh; rather, it was a force that comprehended them both, as if they were but the matter of love, its specific substance. To a woman or to a poem, it said simply: Look! I am alive.63

Against the Darkness  257 The fact that his realization is once again attenuated through representation— Katherine’s book—suggests that letters assemble the disparate realities of his life into a coherent whole. Moreover, as Stoner struggles during the last stage of his life, images of light and darkness reappear and unite: “in the shadowed corridor a sourceless light seemed to glow and dim, pulsating like the beat of his heart; and his flesh, intimately aware of every move he made, tingled as he stepped forward with deliberate care into the mingled light and dark.”64 Here the light and darkness no longer vie for dominance; the light does not flicker, nor does it dispel the gloom. The elements are yin and yang, just as, in the total vision of the novel, craft meets subject and becomes vision. Like Archer Sloane before him, Stoner possesses both the light and darkness, not as separate, independent forces, but as interdependent forces that birth the aesthetic experience. There is, then, a sharp contrast between the lived experience of the world outside of the aesthetic experience and the experience of the world through the aesthetic experience. Without it, as the novel repeatedly shows us, the world is desolate and cruel, and our reactions to it are rightfully despairing. But, the lived experience within literature offers a path beyond despair. Steve Almond writes, “I consider myself blessed to have found this novel, whose every page reminds me why I write: to learn more about what it means to be human, and to use the words I stumble across—as a writer and a reader—to help me bear the most painful moments of that awareness.”65 Without moving life beyond life and invoking, or engaging with literary, or artistic (in general), experience, one is hard-pressed to find a flickering light against the darkness, another possible element in the novel’s recent far-reaching appeal. Kreider believes “that there is nothing better in this life” than this experience.66 The aesthetic experience is about taking the raw material of life, which cannot communicate to one beyond the senses, and raising it to the level of consciousness. It is, perhaps, an act of faith to bely despair with the artistic impulse, as reader and/or creator, for nothing within material reality lifts the banal to the beautiful except consciousness exerting its powers to make the world “mean.” There is always, then, an indeterminate space between material reality—the cold, quantifiable world devoid of poetry—and the artistic impulse toward joy.

Notes 1 The blurbs, in order of appearance, within The New York Review of Books edition of Stoner: Nick Hornby, D. G. Myers, Daniel Okrent, Simon Winchester, and the South Mississippi Sun Herald. 2 Almond (2014). 3 Habash (2013). 4 Kreider (2013). 5 Snow (1986, pp. 103, 104). 6 Ibid., p. 104. 7 Dickstein (2007). 8 Livatino (2010, p. 422).

258  Phillip E. Mitchell 9 For instance, the International MLA Bibliography currently returns only eleven results for a combination of “John Williams” and “Stoner.” 10 Except for McKenna’s work, most of the published works on Stoner are book reviews, interviews with John Williams, or surveys of his novels. 11 McKenna (2015, p. 195). 12 Ibid., p. 193. 13 Ibid., p. 195. 14 Ibid. 15 Dickstein (2007). Dickstein suggests “Williams’s novels share a simple, resonant, sculptured style, eloquent in its restraints.” 16 Whether one might call art transcendent is another matter altogether. 17 Livatino (2010, p. 417). 18 Williams (2003, p. 3). 19 Kreider (2013). 20 Snow (1986, p. 105). 21 Williams (2003, p. 108). 22 Ibid., p. 13. 23 Ibid., p. 11. 24 Ibid., p. 15. 25 Ibid., p. 15. 26 Ibid., p. 27. 27 Ibid., p. 13. 28 Ibid., p. 16. 29 Ibid. 30 Nietzsche (1976, p. 505). 31 Ibid. p. 506. 32 Williams (2003, p. 69). 33 Ibid., p. 70. 34 Ibid., p. 17. 35 Ibid., p. 156. 36 Ibid., p. 246. 37 Ibid., pp. 30, 31. 38 Ibid., p. 31. 39 Ibid., p. 41. 40 Ibid., p. 42. 41 Ibid., p. 37. 42 Ibid., p. 38. 43 Linda Hutcheon believes Don Quixote is an early form of a self-reflecting literary structure that attempts to “unmask dead conventions by challenging, by mirroring” (Hutcheon 1984, p. 18). 44 Livatino (2010, p. 421). 45 Williams, Stoner, 36. 46 Ibid., p. 41. 47 Ibid., pp. 41, 42. 48 Ibid., p. 234. 49 Ibid., p. 235. 50 Ibid., p. 248. 51 Hutcheon (1984, p. 18). 52 McKenna (2015, p. 191). 53 Williams (2003, p. 194), emphasis mine. 54 Hutcheon (1984, p. 29). 55 Kreider (2013). While Kreider’s use of the word “outside” seems problematic in light of my argument, outside seems to suggest the relationship of the reader to the story, the consciousness that “lights” the text, not a magical, or transcendent

Against the Darkness  259 force. Later, for example, Kreider writes that the novel “endures, illumined from within.” This illumination can take place only through the relationship between the reader and the story. 56 Williams (2003, p. 113). 57 Ibid. 58 Ibid. 59 Ibid. 60 Hutcheon (1984, p. 27). 61 Ibid., p. 30. 62 Smith (2013). 63 Williams (2003, p. 250). 64 Ibid., p. 258. 65 Almond (2009). 66 Kreider (2013).

Works Cited Almond, Steve. 2009. “Lost and Found.” Rumpus. Online at January 26, 2009 http://therumpus.net/2009/01/lostand-found-by-steve-almond/ ———. 2014. “You Should Seriously Read Stoner Right Now.” The New York Times. Online at May 9, 2014 www.nytimes.com/2014/05/11/magazine/you -should-seriously-read-stoner-right-now.html?mcubz=1 Dickstein, Morris. 2007. “The Inner Lives of Men.” The New York Times. Online at June 17, 2007 www.nytimes.com/2007/06/17/books/review/Dickstein-t. html?mcubz=1 Habash, Gabe. 2013. “Stoner Finds Overseas Success.” Publishers Weekly. Online at April 20, 2013 www.publishersweekly.com/pw/by-topic/international/internationalbook-news/article/56913-stoner-finds-overseas-success.html Hutcheon, Linda. 1984. Narcissistic Narrative: The Metafictional Paradox. New York, NY: Methuen. Kreider, Tim. 2013. “The Greatest American Novel You’ve Never Heard Of.” The New Yorker. Online at October 20, 2013 www.newyorker.com/books/ page-turner/the-greatest-american-novel-youve-never-heard-of. Livatino, Mel. 2010. “A Sadness unto Bone: John Williams’s Stoner.” Sewanee Review. 118(3): 417–422. McKenna, Tony. 2015. “John Williams’s Novel Stoner and the Dialectic of the Infinite and Finite.” Art, Literature, and Culture from a Marxist Perspective: 190–196. Nietzsche, Friedrich. 1976. “The Birth of Tragedy,” in Philosophies of Art and Beauty: Selected Readings in Aesthetics from Plato to Heidegger, edited by Albert Hofstadter and Richard Kuhns (pp. 496–554). Chicago: University of Chicago Press. Smith, Drew. 2013. “Famous for Not Being Famous: Enough about Stoner.” The Daily Beast. Online at October 13, 2013 www.thedailybeast.com/ famous-for-not-being-famous-enough-about-stoner Snow, Charles Percy. 1986. “Good Man and Foes.” Denver Quarterly 20(3): 103–106. Williams, John. 2003. Stoner. New York, NY: New York Review of Books.

13 Confidence Without Certainty J. Aaron Simmons

Introduction: “You Want to Major in What?!” Rarely do I hear students saying, “I would love to major in chemistry and go to medical school, but my parents tell me that it is impractical, so I am just going to major in philosophy.” And yet I have had numerous students come to me in tears saying that they would like to major in philosophy but their parents won’t let them. Recently, President Donald Trump suggested that immigration preference should be given to those individuals who have advanced degrees in STEM disciplines, but not in the humanities and the arts. During his own presidential campaign, Marco Rubio claimed that “we need more welders and less (sic) philosophers.” Although Rubio has since changed his view on the matter, his comments remain refletive of a broader social assumption about the “value” (or lack thereof) of philosophy. Similarly, business departments might be thriving by training students to function in a market driven world, but those departments that help students to think about how the world is driven seem not to be doing as well. Indeed, maybe we need a new preferred acronym for those areas that receive such social privilege and political applause. I propose that we use “STEM(B)” since it is often actually a business framework that is operatively justifying the concern for STEM fields in the first place. In a time when science is often dismissed as a liberal conspiracy, it is dangerous even to give the impression that the sciences are problematic, and I don’t want to do that here. Rather, as Brandon Inabinet and I have argued elsewhere,1 and David Scott and I have also suggested in relation to the task of the university,2 we must not pit the humanities and the sciences against each other, but instead find ways for the sciences to recognize that they are made better when they are not assumed to be totalizing. Science must avoid scientism if it is to avoid falling into the very thing that it is claimed to be by its despisers. However, perhaps the preference for STEM(B) over the humanities and the arts makes sense. Indeed, according to a popular cultural conception, where science provides answers, philosophy seems to stagnate in questions. Where technology offers determinate progress for human

Confidence Without Certainty  261 society, philosophy tarries with the indeterminacy of existence. While business creates jobs, philosophy creates frustration. In light of such prominent assumptions, however, philosophers are often so worried about remaining relevant that they do little to help matters. Even if it is true that philosophy can help you in business, or that philosophy is likely to up your chances to get into medical school, etc., to turn to such strategies is already to have sold the field upon which one is playing to those who have decided to play a different game altogether. Rather than defend philosophy by making it look more like a science, we need better to articulate what a philosophical life is all about. Accordingly, the question that guides my thinking here is this: how are philosophers to live with confidence in truth-seeking, while lacking certainty about what the truth is? Or, said a bit more technically, how are we to live with determination while also embracing the indeterminacy of life? Philosophy is hard, but so is life. Sometimes, philosophers can quite regrettably make things more difficult than they need to be, but in this case, we must be careful not to do so. Considering indeterminacy should not be a cause for making philosophy even harder, or our language more opaque. Rather, it should be a moment that provides an opportunity to reflect on the relation of philosophy to life in terms that connect it to the lives that we lead. In other words, indeterminacy, questions, and uncertainty are relevant to philosophy because philosophers are existing individuals and life is risky. Yet, the simplicity of this point should not mask the difficult existential task that it presents to us. There are reasons that Friedrich Nietzsche speaks about mountain-climbing as a metaphor for philosophy: the rocks are jagged, the slopes are steep, and the air is thin. And yet, it is also up the mountain that we can begin to see farther, attend carefully to each step we take, and learn to breathe deeply without hyperventilating. We all face the pressures of socio-political, parental, ecclesial, and economic encouragement to seek certainty, find answers, overcome indeterminacy, and supposedly only then to obtain happiness as a result. In response to such a cultural tide, philosophers must demonstrate that recognizing the realities of indeterminacy that attend our epistemic commitments, as well as the questions that are intrinsic to our existential condition, do not lead to a wishy-washy life of compromise, but instead can invite us to an even more compelling manifestation of determination, albeit of a virtuously humbled sort, in our beliefs and actions. My basic thesis, then, is quite simple: confidence can be had without certainty when we distinguish between existential determination and epistemic determinacy. Although the latter is a noble goal in many ways (and sometimes even necessary given particular tasks), it can often become an idol of our own making, such that philosophy begins to become the thing that it ought to oppose: an objective algorithm that reduces morality, faith, justice, truth, and knowledge to a cold articulation of “what all reasonable people ought to believe.”

262  J. Aaron Simmons When philosophy attempts to defend its relevance on the terms of those who would argue for its demise, then the thing being defended is no longer rightly viewed as “philosophy,” but instead merely as another turn in the cultural assumptions underwritten by a corporate logic that reinforces scientistic objectivism and an attitude of certainty. In a world where STEM(B) is assumed to be a default setting for obtaining knowledge and finding success, we might indeed find all sorts of answers, but we would be at risk of no longer asking questions that matter. Were this to be the case, as Henry David Thoreau puts it, there would then likely be “professors of philosophy, but no philosophers.”3 In what follows, I will begin by offering something of a personal engagement with my own identity as a philosopher, and as a teacher, in order to show how the longing for certainty can function as an obstacle to genuine philosophical existence. Then, in the attempt to challenge the logic of objective certainty that I take to underlie STEM(B), I will turn to Michel Henry’s account of barbarism as a genuine threat to philosophy, and life itself. Henry warns of the high costs that accompany allowing epistemic determinacy to replace existential determination. Having set up the contemporary problem via Henry, I will propose a possible response to it by sketching an account of existential determination drawn from Søren Kierkegaard’s discourse, “The Lily in the Field and the Bird of the Air.” What we will see in Kierkegaard, and I will attempt to expand to philosophical living more broadly, is that trust, hope, and risk are definitive of existence in ways that require us to stand firmly while also recognizing that there are always other places that one could (legitimately) stand. Philosophy Makes Me Tired Every semester, I am met with the same question from at least some of my students: “How can you live with so many questions and so few answers?” The motivations for this question are varied, but the two most common that my students provide are that (1) philosophy seems exhausting (and usually this includes a comment about my being hard to live with and expressions of sympathy for my wife!), and/or (2) philosophy seems to eliminate the possibility of ever coming down on a particular side of an issue (interestingly, this is usually the rationale given to me by my more-religious students who are especially invested in the need for certainty of belief about God and morality). Straightaway, let me be clear about one thing: my students are not irrational to hold such views and express such concerns. Their question should trouble all philosophers because it is entirely reasonable within the current hermeneutic framework in which my students attempt to figure out their lives, careers, and identities. I have often said that those who spend the most time considering existence are usually the least good at living well. So, perhaps my students are on to something about the dangers of philosophical pursuits.

Confidence Without Certainty  263 Philosophers are often the punch line of some joke that hints at the uselessness of studying philosophy due to the common view that the “real world” requires something more than merely rigorously questioning it. Unfortunately, philosophers do not do enough to challenge the idea that philosophy is mainly about learning to sit on fences without falling off, rather than living with purpose on the ground. And, in that context, since most folks are very firmly entrenched on one side of the fence or other, maybe philosophy is, as things stand now, actively counterproductive in contemporary society. As any student of global politics will quickly realize (this is especially true in an American context), facility with conceptual nuance and a commitment to linguistic precision will usually spell disaster for one’s ability to connect with the masses. Dumb it down. Make the dichotomies starkly opposed. Keep the choice simple. Eliminate ambiguity. Make those with whom you disagree seem either irrational or immoral. These are often the keys not only to political office, but also to that which frequently serves as the primary justification for a college education in the first place: Success! Such success is then often cashed out as having the determination to overcome obstacles while holding fast to a particular truth . . . about economics, politics, morality, religion, etc. It is those who are troubled by objections who fail to make an impact, or so most of my students seem to think—and in this way, they reflect a general social assumption about not only the uselessness of philosophy, but also about what really matters in life itself, such that philosophy is useless. In the face of social pressures to live with determination in such a way as to minimize the self-doubt that will almost always accompany serious thinking, philosophy might rightly seem to be an obstacle to living well . . . or at least to living richly. Surely, courage in the face of challenges and determination in the face of opposition are virtues to be cultivated, but far too often those who exemplify such virtues do so not by cultivating the life of the mind in the context of persistent questions, but instead by minding the life that they want to achieve, no matter the costs. Here the virtue of philosophical determination runs into the vice of determinately avoiding objections that could undermine your personal confidence and your position within your social group. If you want to go unchallenged, avoid those who will offer critique. If you want to ensure that you will hit your target audience, only speak to audiences that already agree with you. In many ways, then, philosophy runs counter to all of this “real world” advice. Well, with that said, and with all due apologies to the many university marketing staff and admissions officers who claim that it is “real world experience” that allows liberal arts education to remain relevant in the twenty-first century, I contend that philosophers must loudly announce that this is a world that should no longer be allowed to be real. Philosophy is classically defined by the commitment to loving truth, even when that truth is inconvenient, upsetting, or humbling. For the majority of the history of philosophy, the very idea of living well is, itself, defined by

264  J. Aaron Simmons the commitment, as Socrates says, to “follow your beloved wherever it may lead.” Loving truth is, according to philosophy, the basic characteristic of any life that would be “worth living.” Importantly, philosophy is the “love of” wisdom and truth, not the obtainment of such. Philosophy focuses on the constancy of internal transformation in light of truth, what Socrates might term the never-ending investment in soul-leading, rather than on grasping at fleeting externals deemed to be valuable by late market capitalism. As Nietzsche suggests throughout his authorship, and in this sense he is characteristic of the history that he so often eschews, to love truth is to be a seeker, a questioner, and a metaphorical mountain-climber—not a finder, an answerer, or someone who is comfortable on flat land. To love something is to long for it, and then to keep moving. This does not threaten commitment, but instead serves to deepen it. Holding fast to the love of truth is essentially to have faith that truth matters. Faith is best understood as simply risk with direction. Yet, there is nothing simple about faith. To love truth is to be committed to the idea that the pursuit of truth is worth it. Thus, philosophy offers a profoundly different account of the “real world” than in the marketing brochures of all the colleges and universities claiming to prepare students for it. This philosophical account is not a matter of obtaining a job, but instead about being the sort of person who understands why jobs matter, why work is important without being all-consuming, and how devotion to something does not mean no longer being receptive to harsh criticism. Maybe Socrates was right that questioning and examination are the conditions of meaningful life, not obstacles to it. Unfortunately, rather than loving truth, come what may, far too often, living well amounts to naming as true whatever one loves. It is this reality that, I believe, explains my students’ genuine perplexity at the philosophical way of life. Wouldn’t it be a better idea, they say, to strive for what matters in life, which they always take to be obvious and predetermined according to a corporatist theory of value, rather than constantly questioning what, in life, should matter? I admit that the older that I get (currently, at the time of this essay’s publication, I am 40), I am increasingly tempted (or maybe seduced?) by this line of thinking. Importantly, it is not simply propounded by college students and political advisors, but also by a large portion of culture at large— especially within the scientific community, which continues to ask why philosophy should continue to be relevant when it seems like philosophers have made so very little progress over the past few thousand years. If the questions are basically the same for us now as they were for Plato and Aristotle, then perhaps something is wrong with the questions themselves? Or, even more likely, maybe there is something mistaken about the method of the questioners? Although a more sophisticated methodological (and maybe even ontological) critique, this scientific view is reflective of the culture that generates the heart-felt concern from my students. It is not that such questions produce cultural norms, but instead that they reflect the ones already so deeply ensconced that they are simply taken as obviously “true.”

Confidence Without Certainty  265 We all, then, find ourselves faced with two different accounts of the real world. When Morpheus welcomed Neo to the “desert of the real” in The Matrix, it seemed all too tempting to think that he had hit a foundation that allowed for direct access to the way things really are. If only. Instead, we are caught between two narratives—one that operates according to an economic logic of answers in light of the determinacy offered by a context in which STEM(B) is exclusively worth our time and effort, and another that operates according to a philosophical logic of questions in light of the indeterminacy of existence. We might say that on the first model, profit presses upon us, while on the second model we are pressed by the question of whether profits profit us very much at all. Likewise, on the former account, philosophy is a job that seems to serve no real social function. On the latter account, alternatively, philosophy is not a matter of professional activity, but of committed (albeit risky) examined life.4 Yet, unlike Neo, there is no pill that we can take (whether red or blue) that allows us to get out of the narrative in which we find ourselves to decide which version is correct. We are already engaged and can only reflect on the shortcomings of where we find ourselves by owning up to where it is that we currently are. In this way, we are like characters in a story trying to figure out how to become the author of it. If these two narratives were on equal footing, then that would be one thing, but they are not. The former narrative has become an orthodoxy of contemporary social life. According to the dominant narrative, questions are not valued as much as answers, the value of determinacy outstrips that of indeterminacy, and certainty is necessary in order to have confidence. On this account, philosophers are not simply minor characters, but rather are at risk of being written out of the story altogether. Many of the questions asked by philosophers rarely get finally answered, because they are questions that hit at the heart of the human condition. They are not issues to be settled so that a patent can be issued and profits obtained. They are questions about what it means to be the beings that we are. So long as we continue to be such beings, the questions will continue to press upon us—each generation having to ask and answer them anew. As Kierkegaard says, what is truly essential cannot be learned secondhand from history,5 and we might add that it also cannot be made into an objective outcome of a repeatable experiment. What is essential must be learned for oneself as a result of a full commitment of one’s own selfhood. Unless we are at stake in our questions, unless we risk ourselves in asking them, then we are never fully present to ourselves in our own thinking, believing, and acting. Indeed, as Thoreau worried, in such a situation we are likely to come to the end of our lives and realize that we have not spent much time living. So how, then, might philosophers better make the case for having existential determination while still embracing the indeterminacy of existence? In what ways can they make a case for the narrative in which Socrates emerges as at least as worth emulating as Steve Jobs? How can lived confidence occur without the need for epistemic certainty? Here the two general rationales

266  J. Aaron Simmons that underlie my students’ questions about philosophy return with urgency. In the first case, the exhaustion they fear can be framed as a general underlying commitment to the supremacy of objectivist science when it comes to knowledge acquisition. In other words, I think that they think philosophy is just too tiring because they generally assume that only when questions can be answered in fairly short order are they questions worth asking. In the second case, the need for certainty and determinacy assumes a model of existence that ultimately understands morality and religion in objective ways and, thus, eliminates the essential subjectivity that forms the core of life itself. In order to contest these assumptions and directly address their questions in ways that are likely to speak to them where they are (rather than where we wish they were), we must challenge the scientistic and objectivist frameworks that support them. I will turn to Michel Henry to address the first assumption and then to Kierkegaard to address the second assumption. Henry will allow us to see the basic problem and Kierkegaard will hopefully offer potential resources for responding to it. “I Should Have Gone To Medical School”: Henry and the Risk of Barbarism Michel Henry is not a big fan of science . . . or so it can quickly seem. In his book Barbarism (originally published in French in 1987), Henry offers one of the most substantive critiques of modern scientific culture offered in recent philosophical literature. His basic claim is that “we are faced with something that has never really been seen before: the explosion of science and the destruction of the human being.”6 Without going too far afield into a more substantive engagement with his thought, which I have tried to do elsewhere,7 let me provide just a sketch of why he is so worried that the continued rise of science will amount to the death of culture and the eradication of life itself. By thinking about such things with Henry, we will be able to see a possible reason for why the supposed need for certainty is so deeply ingrained in our social consciousness. Almost all of Henry’s authorship concerns what he terms life. By “life,” he is not referencing the object of study of biologists, but instead what might be called the subjectivity that underwrites anyone’s historical existence.8 In this sense, life is necessarily not an object, in the sense of a particular thing in the world alongside other things, and so it is unavailable for objective analysis. It is, rather, the animating depth dimension that allows things to be meaningful. Henry suggests that religion, art, and ethics are three areas, in particular, where life signifies most prominently. Unfortunately, since the rise of modern science, Henry thinks, a concern for life has systematically been occluded by the totalizing belief that knowledge is reducible to what results from scientific processes—and that subjectivity only matters so long as it can be studied in an objective manner. When life becomes a data point, however, it ceases to be life. Of course, stuff can still happen, but none of

Confidence Without Certainty  267 that stuff would be a matter of dynamic interiority—that is, the upsurge of life—but merely amount to pure static exteriority as articulated in the objective discourses of the social sciences.9 Barbarism results when we lose contact with life due to a framework that forces us outside of ourselves in order to know anything at all. Science, then, is not simply a way of knowing the world, but in the context of scientism becomes instead the very arbiter of what world is knowable. Henry basically offers a phenomenological account of why turning people into units of measure amounts to eliminating the very thing that we were attempting to study. Importantly, though, Henry does not mean to suggest that science, as such, is problematic. Instead, his worry is not that people engage in scientific activities, but that science becomes exhaustive for which activities are considered meaningful. For example, we know lots of things about the world due to modern science, but when we think that the only things that are knowable must be knowable according to objectivist methods, the world itself is problematically narrowed. There are drastic metaphysical consequences to this reductionist epistemology. Henry, thus, presents a stark dichotomy here between “truth of the world” and the “truth of life.” Even if we might take issue with this radical opposition, when meeting with parents about their children’s college majors, I continue to realize that they agree with the oppositional framing that Henry articulates. For example, I have heard all of the following claims just in the last few months: “If you are going to major in philosophy, then at least double major in something practical.” “If we are going to recruit majors in philosophy, then we need to show them how this translates into the ‘real world.’ ” “Why would you major in philosophy? Don’t you care about getting a job?” Notice the dichotomy that underwrites these claims: there is a world that matters, and then this other stuff that you can do in college, such as philosophy, that just doesn’t. When pushed, this dichotomy frequently gets developed as a divide between the humanities and STEM(B). Henry’s framework is an extremely helpful analysis of how plausibly to make sense of what is going on in this divide that plays out every day on our campuses, in our political discourse, and in our economic decisions. Accordingly, this can be seen in my students’ expression of exhaustion at the prospect of doing philosophy for any length of time. On an objective model, questions are designed such that they can be answered. According to a model that recognizes the ineradicable subjective dimension of life, questions emerge because life, itself, is messy. For many students, the idea of being fully invested in a life of questioning and truth-seeking has become replaced by the need to find ourselves as already in possession of the answers and defined by particular truths. Sure, experimentation continues and new questions get asked, but only insofar as nothing essential changes because what constitutes our essence is never understood to be at stake. So long as we think that we must justify the truth of life in terms of the truth of the world, life itself will never be justifiable because it is unable to

268  J. Aaron Simmons be countenanced in those terms. It is for this reason that philosophers will never be able to make a case for their relevance according to a narrative that narrates them out of the story. In such a situation, what is most fundamental becomes the most dispensable thing of all. Henry explains this by looking at the experience of a biology student: Let us consider a biology student who is reading a work about the genetic code. . . . The knowledge contained in the biology manual that was assimilated by the student during her reading is scientific knowledge. The reading of the book uses a knowledge of consciousness; it consists partly in the perception of words, that is to say, in the sensible intuition of marks written on paper, and partly in the intellectual grasp of the ideal meanings that the words carry. Together, these meanings form the sense of the book, that is, the scientific knowledge contained in it. The knowledge that made possible the movements of the hands and the eyes, the act of getting up, climbing the stairs, drinking and eating, and resting is the knowledge of life. If one were to ask which of these three types of knowledge is fundamental, it would be necessary to reject the prejudices of our time all at once: the beliefs that scientific knowledge is not only the most important but in reality the only true knowledge; that knowledge means science, that is, the type of mathematical knowledge of nature introduced at the time of Galileo; that everything prior to this arrival of rigorous science in the West was only a mass of disordered knowledge and confused feelings, if not prejudices and illusions. One should not forget, however, that beginnings are always what is the most difficult.10 Henry’s point is simple: what we take for granted as the most important thing, and even as the only important thing, is actually dependent on that which we have systematically ignored, devalued, and eliminated. That is, subjectivity is the basis for any objective concern. Religion, art, and ethics are the domains in which meaning gets worked out such that science can then affirm itself as meaningful. Notice that Henry is not saying that the biology student shouldn’t study genetics. Instead, he is pointing out that in order to study genetics, we are already deploying a knowledge that far too often STEM(B) disciplines dismiss when they become totalizing. In so doing, a society entirely devoted to STEM(B) doesn’t just eliminate the humanities, but even more interestingly eliminates itself. The death of culture is not the triumph of the sciences, it is the death of an historical situation in which science could be conducted as a meaningful human practice. When my students are perplexed by my devotion to a life of questioning because they consider it to be exhausting, I find that they are reflecting the assumption that serves to reinforce the scientistic objectivism about which Henry worries. They are more interested in the end of things (i.e., the answer), than they are about the beginning of things (i.e., the question that

Confidence Without Certainty  269 motivates a life searching for such an answer). But, as Henry reminds us, we must not forget “that beginnings are always what is the most difficult.” In an age of STEM(B), the beginnings are taken for granted as passé and old news. The new, the innovative, the next-thing, are always where the interest lies. Given this framework, it makes sense that students would ask: “Why should I study philosophy if I want to go to medical school?” Yet, this question throws us back upon the distinction between existential determination and epistemic determinacy. “To know” is not the same thing as “to be.” And, knowing about being is not the same thing as being a knower. If the goal of one’s education is to get a job, then perhaps one should just do what it takes to get to medical school, but if the point of jobs is to engage living bodies in meaningful activities whereby culture can flourish, then we must be willing to risk ourselves by being existentially committed even when we lack epistemic certainty. Even if we help our students to get high-paying jobs, unless we have enabled them to understand what it means to be invested in the messy and uncertain task of living, then we have failed them. Toward the end of Barbarism, Henry himself takes up this question of what it means to prepare our students not only for employment, but for life: Preparing an individual for a job and for entering into society can take on very different senses, depending on whether or not it is a question of a completely economic society obeying an economic or technological teleology. Economic participation, for example, presupposes the actualization of some potentialities of the individual, while it is indifferent to his or her general development.11 Henry goes on to note that the general slow pace at which universities produce knowledge is what allowed them to be culture shapers, rather than merely ideology reflectors.12 Inviting our students to slowness and intentional living does not mean that we are ignoring the practical realities of a market economy, it just means that we are refusing to allow our students to become nothing more than units of economic value. Helping them see that existential determination is not threatened by the indeterminacy of a realm of subjectivity irreducible to objective inquiry is essential to providing them the skills necessary to cope with their own finitude—regardless of what job they end up holding. “Life is Not Simply About a Job”: Kierkegaard on Silence, Obedience, and Joy Whenever I feel disconnected from humanity, I turn to Kierkegaard and there, finding God and losing myself, I rediscover the faces of others. This is how I once described why I keep reading Kierkegaard even though he is so often troubling to everything that I otherwise hold dear (e.g., God, family,

270  J. Aaron Simmons morality, democracy, etc.). Indeed, Kierkegaard doesn’t allow anyone to rest content in themselves, in their God, or in their rationalism. Egoism, selfcertainty, moral clarity, and religious absolutes are all disturbed in an account that I have summarized elsewhere with the phrase “God is trouble.”13 As with Henry, Kierkegaard’s authorship is so complex that any attempt to offer a detailed analysis here would be merely a distraction. However, I do find that his work offers profound resources for being able to respond to the situation that Henry has outlined, while also allowing us better to understand the rationale behind my (especially religious) students’ worry that embracing uncertainty amounts to being unable to stand anywhere with confidence. There are many Kierkegaardian texts upon which we could draw here, but I will focus simply on his little Upbuilding Discourse from 1849 entitled, “The Lily in the Field and the Bird of the Air.”14 Kierkegaard begins the discourse with a short prayer in which he asks his “Father in heaven” to help him understand “what it is to be a human being” in light of the difficulties presented by the “crowd” such that the “company of people” causes us to forget this existential identity.15 By turning to the example of the lily and the bird, Kierkegaard suggests that the key to subjectivity is found in learning to be silent, obedient, and joyful. These might seem like an odd trinity of qualities, especially in light of a world in which brashness, disruption, and anxiety tend to be all too common. However, Kierkegaard’s explicitly religious account of human existence can be appropriated more broadly as a model of what confidence can look like without certainty. For example, when he cites the gospel passage “Seek first God’s kingdom and his righteousness,”16 we ought to hear him inviting us to reflect on what it means to get our priorities straight as finite beings. As Henry said, the beginnings are often the most difficult, but the most important. Similarly, Kierkegaard contends that learning to listen, to be open, to be exposed, is prior to, and more important than, speaking, being closed off, and narrowly self-protective. In some ways mirroring Henry’s distinction between the truth of the world and the truth of life, Kierkegaard contrasts being “out there with the lily and the bird” with “speaking and conversing with other human beings.”17 He is not being misanthropic here, but simply reminding us of the way in which social assumptions tend to function as mechanisms of hermeneutic closure, rather than invitations to existential exploration. “Out there with the lily and the bird,” Kierkegaard says, “you are aware that you are before God.”18 This awareness, which comes when we cultivate a genuine receptivity—the capacity to be silent and listen—amounts to a radical humility before that which overwhelms our epistemic capacity, our scientific methodology, and our political conceptions. Kierkegaard was determinately a Christian, but in order to get his point, one does not have to adopt his theological commitments. Humility is a virtue whether or not one locates it in relation to religion. But, it is a virtue that runs counter to the logic of scientism. The humility that Kierkegaard

Confidence Without Certainty  271 recommends is more than merely an epistemic realization that one might be wrong about a particular belief. Instead, it is an existential recognition that what really matters about life might not be knowable in standard ways: “Would that you in silence might forget yourself, what you yourself are called, your own name.”19 Notice that instead of claiming that with the lily and the bird one “knows that you are before God,” he says “you are aware that you are before God.” Being aware of something announces the sort of receptivity that Henry locates as activated in the culture-producing practices of life. The humility cultivated in silence does not yield hesitancy and self-doubt, however. Instead, Kierkegaard boldly claims that it requires obedience and yields joy. For Kierkegaard, this obedience is specifically in relation to God’s will, but in the broader non-theological context in which we are working here, we can read him as suggesting that we must make existential decisions where we are, even in the face of that which outstrips our comprehension and cognitive grasp. “What marvelous security,” he writes, “in finding the unconditioned and in having one’s life in it!”20 All of us will come down somewhere on that which we take to be unconditioned—i.e., that which outstrips all pretense to epistemic determinacy. Kierkegaard’s point is that when we allow that unconditioned to be manifest in concert with radical existential humility, we are able to stand determinately while still admitting of indeterminacy. When taking oneself up in this way, certainty becomes more of an idol of which to be wary, than a goal toward which to strive. As Kierkegaard explains: If, when the moment comes to fly away, the bird is ever so certain that its present situation is good just as it is, so that by flying away it will let go of the certain in order to grasp the uncertain—yet the obedient bird immediately starts out on its journey. Simply, with the help of unconditional obedience, it understands only one thing but understands it unconditionally—that now is the unconditional moment.21 The point is not that we ultimately eliminate all doubt and obtain objective evidence that allows us brashly to affirm ourselves, our own community, our own perspectives, as the most valuable, but that we humbly articulate ourselves as fully invested, here and now, even though we may need to move on tomorrow. Like Abraham, who is said to have gotten up early in the morning in order to make the journey to Mt. Moriah, here we see the bird “immediately starts out.” Without delay, they both move in obedience— realizing the risk, but also investing themselves by assuming the posture of childlike trust. Following the biblical recommendation, Kierkegaard views the high point of maturity to be the ability to trust like a child. Lacking objective certainty in a relation of trust does not amount to hesitation in belief or action—and this is something that my students often fail to appreciate (especially my religious students). For many of them, if there

272  J. Aaron Simmons is even a moment of question when it comes to God, career, politics, profit, etc., then they are shaken to their core. Kierkegaard demonstrates, alternatively, that in the face of the tremendous mystery of the unconditioned that the lily and the bird become models of peace. They have joy because their joy is not a matter of obtaining external things, but of articulating themselves as properly formed internally in humility and obedience. “What is joy, or what is it to be joyful?” Kierkegaard then asks.22 His answer is striking: “It is truly to be present to oneself; but truly to be present to oneself is this today, this to be today, truly to be today.”23 Joy is not found in getting something tomorrow that will allow you to be happy, but instead in finding oneself happy today in the very act of striving toward selfhood. This joy is what my students often sidestep in their practical concerns with the “real world” that they have been handed. Specifically, they come to college in order to get out of college. They get a job in order to get another one. Their fear of uncertainty causes them to fear losing themselves. Kierkegaard helps us to understand how to respond to their fears by overturning the hermeneutic framework according to which such fears are licensed. Look, I do not care whether folks major in philosophy as opposed to business or chemistry. Rather, my hope is that business leaders, chemists, and philosophers are all defined by living on purpose while recognizing that there is nothing obvious about how best to live. Displaying existential determination is not about being cocky, but about having confidence due to one’s being appropriately humble. Appreciating epistemic indeterminacy is not a matter of just lacking knowledge, but of realizing that if we embrace a life defined by trust, we are likely to become aware of far more than we are able to know. Conclusion: Simone Weil and the Remains of Hope As a way of bringing this chapter to a close, let me suggest that reading Simone Weil is always a good idea. She is not someone who had a taste for certainty, but her existential determination is intimidating. Her authorship reflects a passionate embrace of indeterminacy as a way of life, but importantly it would be difficult to find someone who lived with more intentionality. Living on purpose, she faced affliction with joy, but never with naïveté or blindness. The struggle of living animates her work, and the anguish that defines that struggle is never far from her account of the love of God and the importance of loving others. Yet, despite that, she continued to live . . . on purpose . . . in joy. Perhaps the best summary of her particular manifestation of existential determination despite lacking epistemic determinacy can be found in her description of genuine understanding: Man only escapes from the laws of this world in lightning flashes. Instants when everything stands still, instants of contemplation, of pure intuition, of mental void, of acceptance of the moral void. It is through such instants that he is capable of the supernatural. Whoever endures a

Confidence Without Certainty  273 moment of the void either receives the supernatural bread or falls. It is a terrible risk, but one that must be run—even during the instant when hope fails. But we must not throw ourselves into it.24 Weil’s words here are reminiscent of Emmanuel Levinas’s claim that moral life is “a fine risk to be run.”25 The idea shared by Henry, Kierkegaard, Weil, and Levinas is that doing what matters is never a matter of doing what is obvious and certain. Existential determination is hard work. Living on purpose comes with costs. This is why Weil notes that “we must not throw ourselves into it.” Being open to indeterminacy must neither be confused with losing ourselves in a mealy-mouthed version of relativistic nihilism, on the one hand, nor with rash impulsiveness, on the other hand. It takes courage and humility to “run” this “terrible risk,” but it also requires wisdom to know when to stand where you are. As with Kierkegaard’s bird, sometimes the real risk is that we choose not to move, but this never means that we are no longer moved by the arguments otherwise. It just means that following truth wherever it may lead eventually will require us to make a decision and then to live in light of it. So, let me close by returning to my students’ questions, to their parents’ opposition, to our politicians’ consternation, and to our market logic that often only leaves room for STEM(B) to be valued. I have tried very hard not to dismiss any of these responses as silly. When we take all of these challenges to philosophical existence as challenges worth meeting head-on, we are able to own up to the realities that shape their starting assumptions. My suggestion to philosophers wanting to make an impact on those who begin by thinking philosophy is a waste of time is that we start by interrogating the theory of value by which they would deem such things worthless. Don’t try to work within their assumptions, but make their assumptions clear to them in order that they can own up to them, rather than simply allowing them to function as obvious truths about the “real world.” Not very much is obvious, but it requires a lot of work to get folks to see that. Accordingly, philosophers should be those people who take seriously what everyone else takes for granted. Don’t prepare them for the “real world,” but instead invite them to consider what world they want to make real! I start every one of my introduction to philosophy courses with David Foster Wallace’s This Is Water.26 Therein, he reflects on the egoism that functions as the “default setting” for how we engage the world. Wallace does not offer a moralistic account of how to live, but instead invites his readers to reflect on what counts in living. In order to motivate such reflection, Wallace does not offer grand stories of moral saints, but instead provides anecdotes about standing in line at the grocery store and getting stuck in traffic. His point is not that we should live in this way or that way, or even that we should abandon our default settings, but simply that we should own up to having them in order that we finally take seriously the questions that attend living in any way whatsoever. As he explains, the important thing

274  J. Aaron Simmons about education is not that we learn how to think, but rather that we get to choose what to think about. Choice is real, and it comes with risks. As concerns existence, not choosing is not an option, neither is playing it safe. Ultimately, then, Henry’s worry about the emergence of barbarism must be taken seriously, but we must not be overwhelmed by it. Instead, as we saw in Kierkegaard, with risk comes trust. Such trust does not discharge our existential determination to work for a future worth inhabiting because we are sure things will work out one way or another—to do this would simply enable another mode in which barbarism can continue to expand unabated. Rather, trust occurs at the intersection of risk and hope—hence, the joy of the lily and the bird. We very well might “fall,” as Weil rightly notes, rather than receiving that “supernatural bread” for which we long. But, here we stand, hopefully. I am not certain about very much, but I am confident that a whole lot is worth hoping for. Indeterminacy remains because life continues. “In the end there is the unmanageable,” as David Kangas once told me in personal correspondence, but were it otherwise, were the unmanageable to be tamed and fit into our boxes that are neatly labeled according to the STEM(B) categorizing system, then the loss would not simply be that higher education would become less compelling, but that the human condition would become unrecognizable. Without indeterminacy, without the unmanageable, without the questions that continue, without the constant search for the supernatural bread, without the risk that we fall before receiving it, Kangas rightly says, “we would be trivial beings.” So, I stand. On purpose. Here and now. With determination, but not without risk. With hope because of the trust that defines my life and my identity. In the end, hope remains because although indeterminacy continues . . . so do we—and that matters, which is, perhaps, enough for now.

Notes 1 Simmons and Inabinet (2018). 2 Simmons and Scott (2017). 3 Hadot (2005). 4 Hadot (1995). 5 Kierkegaard (1985). 6 Henry (2012, p. 3). 7 See Simmons and Benson (2013) and Simmons and Scott (2017). 8 Henry (2012, p. 6). 9 Ibid., p. 18. 10 Ibid., p. 11. 11 Ibid., p. 122.

Confidence Without Certainty  275 2 Ibid., pp. 123–124. 1 13 See Simmons (2016). 14 Kierkegaard (1997, pp. 2–45). 15 Ibid., p. 3. 16 Ibid., p. 10. 17 Ibid., p. 17. 18 Ibid., p. 17. 19 Ibid., pp. 18–19. 20 Ibid., p. 26. 21 Ibid., p. 29. 22 Ibid., p. 39. 23 Ibid., p. 39. 24 Weil (1952, p. 11). 25 Levinas (1997, p. 120). 26 Wallace (2009).

Works Cited Hadot, Pierre. 1995. Philosophy as a Way of Life, edited by Arnold I. Davidson. Malden, MA: Blackwell. Hadot, Pierre. 2005. “There are Nowadays Professors of Philosophy, but not Philosophers,” translated by J. Aaron Simmons, with notes by Mason Marshall. The Journal of Speculative Philosophy. 19(3)(Fall): 229–237. Henry, Michel. 2012. Barbarism, translated by Scott Davidson. London: Continuum. Kierkegaard, Søren. 1985. Philosophical Fragments/Johannes Climacus, edited and translated by Howard V. Hong and Edna H. Hong. Princeton: Princeton University Press. ———. 1997. Without Authority, edited and translated by Howard V. Hong and Edna H. Hong. Princeton: Princeton University Press. Levinas, Emmanuel. 1997. Otherwise than Being, Or Beyond Essence, translated by Alphonso Lingis. Pittsburgh, PA: Duquesne University Press. Simmons, J. Aaron. 2016. “Personally Speaking . . . Kierkegaardian Postmodernism and the Messiness of Religious Existence.” International Journal of Philosophical Studies. 24(5): 685–703. Simmons, J. Aaron and Brandon Inabinet. 2018. “Retooling the Discourse of Objectivity: Epistemic Postmodernism as Shared Public Life.” Public Culture. 30(2): 221–243. Simmons, J. Aaron and Bruce Ellis Benson. 2013. The New Phenomenology: A Philosophical Introduction. London: Bloomsbury. Simmons, J. Aaron and David Scott. 2017. “Is there Life after Barbarism?: Phenomenological Reflections on Science and the Future of the University.” Pli: The Warwick Journal of Philosophy. 28: 1–31. Wallace, David Foster. 2009. This is Water. New York, NY: Little, Brown and Company. Weil, Simone. 1952. Gravity and Grace, translated by Emma Craufurd. London: Routledge and Kegan Paul.

Part IV

Asian Perspectives and Cosmological Concerns

14 Heidegger and Dō gen on the Ineffable Filippo Casati and Graham Priest

Being is the indeterminate immediate; it is free from determinateness in relation to essence and also from any which it can possess within itself. Nothing, pure nothing: it is simply equality with itself, complete emptiness, absence of all determination and content—undifferentiatedness in itself. Hegel, Logic.1

1. Introduction Indeterminacy can mean many things. For example, it can have an epistemic sense. Something is indeterminate in this sense if we cannot determine (at least at present) whether or not it is true. Thus, we might say that it is currently indeterminate whether there is intelligent life anywhere else in the galaxy. Presumably, either there is or there isn’t, but we just don’t know which (yet). Indeterminacy can also have a metaphysical sense. Something is indeterminate in this sense just if there is no fact of the matter. It’s not just that we don’t know whether something is true or false, but reality itself, as it were, leaves the matter undefined. Examples of indeterminacy, in this sense, are bound to be philosophically contentious. However, as just one example: some philosophers think that vague concepts give rise to indeterminacy of this kind.2 Suppose that something changes colour slowly from red to blue. There is an area of indeterminacy in the middle: at some point, it’s not true to say of the object that it is red, and it’s not true to say that it isn’t. The reality of the situation determines no verdict on the matter. An object, then, is indeterminate with respect to some characteristic, F, if it is the kind of thing to be F, but it is neither F nor not-F.3 In our example, the characteristic is red. The ineffable—something about which one can say nothing—if there is such a thing, delivers an extreme case of indeterminacy. It is indeterminate with respect to every characteristic. Since one can say nothing of it, there is

280  Filippo Casati and Graham Priest no characteristic, F, such that it is either F or not-F. For each of these would say something of it. Of course, one might hold that there is nothing which is ineffable. However, many great philosophers or philosophical traditions have held there to be some things that are ineffable. In what follows, two such will concern us. One is Martin Heidegger (1889–1976); the other is the tradition of Zen Buddhism, especially in the form given to it by Dōgen Kigen (道元希玄, 1200–1253). Heidegger probably needs no introduction to readers of this volume. Dōgen is one of the most important Japanese Zen philosophers, who founded the Sōtō School of Zen in Japan. The similarity between certain aspects of Heidegger’s thought and Zen has been noted by many;4 and it is only one of these which will concern us here: their treatment of the ineffable. Both hold that something is ineffable; both have their reasons for this. But it does not require deep thought to see that there is an issue here. The ineffable is completely indeterminate. But any reason as to why something is ineffable, must say something about it, and so attribute some characteristics to it. Indeed, even to say that something transcends all determination is to give it a determination. Thus, for some characteristics, F, the ineffable thing is either F or not-F, even though there is no such characteristic. The contradiction is clear. So clear, indeed, that philosophers who find themselves in the situation of talking about the ineffable—Dōgen and Heidegger included—are well aware of it. Those who pin their colours to the mast of the Principle of Non-Contradiction try to take some evasive action. Such is rarely successful, though this is not the place to go into these matters.5 For this was not the response of either Dōgen or Heidegger (at least, in the case of the latter, in his later thought). Both simply rejected the Principle of Non-Contradiction. It may well apply to some things, but the thing which is ineffable is not one of them. One can, indeed, talk of the ineffable, characterize the uncharacterizable. In Section 2, we will discuss Dōgen’s philosophy. We will look at the Mahāyāna notion of ultimate reality (Section 2.1) and note that, according to this tradition, ultimate reality is ineffable. Hence, speaking about it, as the tradition does, is contradictory (Section 2.2 and 2.3). We will show that Dōgen endorses this contradiction (Section 2.4). Finally, we will see how Dōgen takes the speech of effability and silence of ineffability to be mutually necessitating (Section 2.5). In Section 3, we will discuss Heidegger’s philosophy. We look at his notion of being, its ineffability (Section 3.1), and the contradiction this entails (Section 3.2). We will see how Heidegger comes to accept this contradiction after the Kehre (Section 3.3). Finally, we will show that he, too, takes speech and silence about the matter to be mutually necessitating (Section 3.4).

Heidegger and Dō gen on the Ineffable  281 Finally, in Section 4, we will see that what is at issue as the subject of the contradiction we are concerned with may, for both thinkers, be seen as the same thing: nothingness. This is shown for Heidegger in Sections 4.1 and 4.2, and, with the help of Nishitani, for Dōgen in Sections 4.3 and 4.4.

2. Dō gen 2.1  Ultimate Reality Let us, then, start with Dōgen. Dōgen was a Zen (禪, Chin: Chan) Buddhist philosopher. But Zen is one kind of Mahāyāna Buddhism, and one cannot understand his thought if one does not understand some central aspects of that whole tradition. So before we get to him, we will have to spend some time on this.6 Let us start by seeing why, in Dōgen’s tradition, there is something ineffable. All schools of Buddhist philosophy hold there to be a distinction between conventional and ultimate reality. Conventional reality (Skt: saṃvṛti satya),7 all schools agree, is the reality which we normally experience: our Lebenswelt. There was less agreement among different Buddhist schools about what, exactly, ultimate reality (Skt: paramārtha satya) is; but in Mahāyāna Buddhism, it was generally agreed that conventional reality is a conceptual construction, and that ultimate reality is what remains if one strips off from conventionally experienced things all conceptual overlays. As Nāgārjuna (fl. 1st or 2nd c. CE), the foundational philosopher of Mahāyāna Buddhism (and specifically, its Madhyamaka School) says in his immensely influential Mūlamadhyamakakārikā (Fundamental Verses of the Middle Way, MMK):8 Not dependent on another, peaceful and Not fabricated by mental fabrication, Not thought, without distinction. That is the character of reality. Indeed, so important is the point, that it is made in the dedicatory verses of the text:9 I prostrate to the Perfect Buddha The best of teachers, who taught that Whatever is dependently arisen is Unceasing, unborn, Unannihilated, not permanent, Not coming, not going, Without distinction, without identity, And free from conceptual construction.

282  Filippo Casati and Graham Priest One does not need to unpack here all of what is going on in these quotations. It suffices that Nāgārjuna is talking of ultimate reality—he is obviously not talking about conventional reality!—and both quotations, in their way (‘mental fabrication’, ‘conceptual construction’), say this is concept-free. 2.2 Ineffability Since ultimate reality is concept free, it follows immediately that one cannot say what it is like: it is ineffable. This is not to say that it cannot be experienced; just that the experience cannot be characterised. It is a simple thatness (Skt: tathātā). The Mahāyāna tradition is well aware of the ineffability of ultimate reality. Thus, Nāgārjuna himself points this out:10 The victorious ones have said That emptiness is the elimination of all views. For whomever emptiness is a view That one will accomplish nothing. The views in question are, of course, those concerning ultimate reality. And later we have:11 ‘Empty’ should not be asserted. ‘Non-empty’ should not be asserted. Neither both nor neither should be asserted. These are used only nominally. How can the tetralemma of permanent and impermanent, etc. Be true of the peaceful? How can the tetralemma of finite, infinite, etc. Be true of the peaceful? The topic of these verses is, again, ultimate reality, and they are saying that none of the four kinds of things one can say about it is applicable. Moreover, in endorsing the ineffability of ultimate reality, Nāgārjuna is simply echoing the verse summary of the Aṣṭasāhasrikā Prajñāpāramitā Sūtra (Perfection of Wisdom Sūtra of 8,000 Lines), which says:12 All words for things in use in this world must be left behind, All things produced and made must be transcended— The deathless, the supreme, incomparable gnosis is then won. That is the sense in which we speak of perfect wisdom. But now there is an obvious issue. Ultimate reality is ineffable, but much is said in the tradition about it. In the quotations of the previous

Heidegger and Dō gen on the Ineffable  283 subsection, Nāgārjuna himself says many things. Indeed, even to explain that it is ineffable because it transcends conceptual imposition is to talk about it. It is patent to anyone after a moment’s thought that there is a contradiction here. The ineffable is being spoken of. Unsurprisingly, then, the Mahāyāna tradition is well aware of the matter; and a number of Mahāyāna philosophers take evasive action.13 For example, the Tibetan philosopher Gorampa14 is as clear as his Mahāyāna predecessors that the ultimate is ineffable. He says in his Synopsis of Madhyamaka, 75:15 The scriptures which negate proliferations of the four extremes refer to ultimate truth but not to the conventional, because the ultimate is devoid of conceptual proliferations, and the conventional is endowed with them. But he also realises that he talks about it. Indeed, he does so in this very quote. Gorampa’s response to the situation is to draw a distinction. Kassor describes matters succinctly thus:16 In the Synopsis, Gorampa divides ultimate truth into two: the nominal ultimate (don dam rnam grags pa) and the ultimate truth (don dam bden pa). While the ultimate truth . . . is free from conceptual proliferations, existing beyond the limits of thought, the nominal ultimate is simply a conceptual description of what the ultimate is like. Whenever ordinary persons talk about or conceptualize the ultimate, Gorampa argues that they are actually referring to the nominal ultimate. We cannot think or talk about the actual ultimate truth because it is beyond thoughts and language; any statement or thought about the ultimate is necessarily conceptual, and is, therefore, the nominal ultimate. It does not take long to see that this hardly avoids contradiction. If all talk of the ultimate is about the nominal ultimate, then Gorampa’s own talk of the ultimate is this. And the nominal ultimate is clearly effable. Hence, Gorampa’s own claim that the ultimate is devoid of conceptual proliferations is just self-refuting. This is, hence, no way out of the contradiction: it merely relocates it. 2.3  The Vimalakı¯ rti Nireśa Su¯ tra The evasion of contradiction is one response to talking of the ineffable in the Mahāyāna tradition. Another is simply to embrace it. This comes out clearly in a sūtra called the Vimalakīrti Nireśa Sūtra (Sūtra of the Teachings of Vimalakīrti—who is, in the dialogue, a very astute Buddhist layman from Licchavi). The sūtra is an Indian Mahāyāna text, of uncertain date, but

284  Filippo Casati and Graham Priest possibly about the 1st century CE. One of its central concerns in the overcoming of dualities, of which the duality between effability and ineffability is a central one. At one point in the sūtra, a goddess appears in the room, and causes petals to flutter down. These slide off enlightened people, but stick to people who are unenlightened. The petals stick to Śāriputra (a hero of a number of the pre-Mahāyāna sūtras), and he is not very happy about this. A conversation between him and the goddess ensues:17 Then the venerable Śāriputra said to the goddess, “Goddess, how long have you been in this house?” The goddess replied, “I have been here as long as the elder has been in liberation.” Śāriputra said, “Then, have you been in this house for quite some time?” The goddess said, “Has the elder been in liberation for quite some time?” At that, the elder Śāriputra fell silent. The goddess continued, “Elder, you are ‘foremost of the wise!’ Why do you not speak? Now, when it is your turn, you do not answer the question.” Śāriputra: Since liberation is inexpressible, goddess, I do not know what to say. Śāriputra, appealing to the idea that enlightenment, the realisation of ultimate reality, is ineffable, takes the 5th Amendment. The goddess is not impressed (ibid.): Goddess: All the syllables pronounced by the elder have the nature of liberation. Why? Liberation is neither internal nor external, nor can it be apprehended apart from them. Likewise, syllables are neither internal nor external, nor can they be apprehended anywhere else. Therefore, reverend Śāriputra, do not point to liberation by abandoning speech! Why? The holy liberation is the equality of all things! The reply is dark. The thought would appear to be that words are not something over and above ultimate reality, which can be—indeed, must be— peeled off of it. They are part of it, and so can be used to describe it. But whatever the exact meaning of the goddess’ words, it is clear that she says that one can speak about ultimate reality. If one left the text at this point, one might just think that the doctrine of the ineffability of ultimate reality had been dismissed. But this is not so. Two chapters later there is a chapter entitled ‘Entering the Gate of Non-Dualism’. As the title suggests, the topic of discussion turns explicitly

Heidegger and Dō gen on the Ineffable  285 to the question of what it means to transcend duality, that is, realise the ultimate. Many Bodhisattvas (beings on the path to enlightenment) are brought into the discussion, and each takes it in turn to say what this means. The last Bodhisattva to speak is the most important of them all. This is Mañjuśrī, the Bodhisattva of Wisdom—so he should know what he is talking about:18 Mañjuśrī replied, “Good sirs, you have all spoken well. Nevertheless, all your explanations are themselves dualistic. To know no one teaching, to express nothing, to say nothing, to explain nothing, to announce nothing, to indicate nothing, and to designate nothing—that is the entrance into nonduality.” Then, Vimalakīrti, the real hero of the dialogue, is asked what he thinks (ibid.): Then the crown prince Mañjuśrī said to the Licchavi Vimalakīrti, “We have all given our own teachings, noble sir. Now, may you elucidate the teaching of the entrance into the principle of nonduality!” Thereupon, the Licchavi Vimalakīrti kept his silence, saying nothing at all. The crown prince Mañjuśrī applauded the Licchavi Vimalakīrti: “Excellent! Excellent, noble sir! This is indeed the entrance into the nonduality of the bodhisattvas. Here there is no use for syllables, sounds, and ideas.” Vimalakīrti remains silent. But unlike the silence of Śāriputra, this is praised. What is the difference? The context. The silence of Vimalakīrti acquires its meaning from what Mañjuśrī has just said about transcending duality. (If Vimalakīrti had been silent because he hadn’t heard this, or just plain fallen asleep, it could not have had the same significance.) Mañjuśrī has just said that you cannot speak about the ultimate. (So Mañjuśrī contradicts himself.) Vimalakīrti shows the same thing. The sūtra, then, endorses speaking of the ineffable.19 Although the Vimalakīrti Nirdeśa Sūtra is an Indian text, it actually had little impact on the development of Indian Buddhism, as judged by the Indian Mahāyāna commentarial tradition. It rapidly became a very well known sūtra in East Asia, however, and finds a core place in East Asian Buddhism. 2.4 Dō gen With this background, we may now turn to Dōgen. When Buddhism entered China around the turn of the Common Era, it met the indigenous philosophy

286  Filippo Casati and Graham Priest of Daoism, which was to exert an enormous influence on the development of Chinese Buddhism. By the 6th century of the common era, a number of distinctively Chinese schools of Buddhism had developed. They were all, however, in the Mahāyāna tradition. Chan was one of these. Traditionally, it was taken to have been brought to China by the Indian or Central Asian monk Bodhidharma (Chin: Damo, 達摩), fl. 5th c. CE, who thereby became the first patriarch of Chan, and the 28th of Buddhism itself (the Buddha being the first). Chan has many distinctive features. Perhaps the most notable of these was the development of various techniques designed to induce in the student an experience of ultimate reality, unmediated by any conceptual overlay.20 This is not the place to go into these matters, however.21 Starting in about the 6th century, virtually all of the schools of Chinese Buddhism entered Japan via the Korean Peninsula. Dōgen trained as a Tendai (Chin: Tientai, 天台) Buddhist monk.22 But, according to the traditional story, he became dissatisfied with this and went to China in search of something better. There he met Sōtō (Chin: Caodong, 曹洞); and when he returned to Japan, he founded this school of Zen there. Dōgen gave many lectures to his monks. These are recorded in his Shōbōgenzō (Treasury of the True Dharma Eye, 正法眼藏). The texts are very difficult to decode.23 They refer frequently to other Buddhist texts, play with language, reinterpret standard Buddhist ideas. Sometimes the aim would appear to be as much to disrupt the listener’s thought as to expound Dōgen’s own. For that reason, they are singularly difficult to translate, and translators can come up with things that appear radically different. Having uttered this warning, let us now turn to what Dōgen has to say about the nature of ultimate reality. As much as for any other Mahāyāna philosopher, it is ineffable. Thus, in the lecture of the Shōbōgenzō entitled Hosshō (The Nature of Things, 法性), he says:24 [CP: It is wrong to think that] the nature of things will appear when the whole world we perceive is obliterated, that the nature of things is not the present totality of phenomena. The principle of the nature of things cannot be like this. The totality of phenomena and the nature of things are far beyond any question of sameness or difference, beyond talk of distinction or identity. It is not past, present, or future, not annihilation or eternity, not form, not sensation, not conception, conditioning, or consciousness—therefore it is the nature of things. But he, like his predecessors, it prepared to talk about it. Indeed, he does so in the above quotation. What does he make of this? Note, first, that Dōgen often appears happy to endorse contradictions. For example, in the Shōbōgenzō lecture entitled Shoji (Life and Death, 生 死), we find:25

Heidegger and Dō gen on the Ineffable  287 Just understand that birth-and-death is itself nirvāṇa. There is nothing such as birth and death to be avoided. There is nothing such as nirvāṇa to be sought. Only when you realize this are you free from birth and death. And in the lecture Genjō kōan (The Issue at Hand, 現成公案), we find:26 As all things are Buddha-dharma, there is delusion and realization, practice, birth and death, and there are Buddhas and sentient beings. as the myriad things are without an abiding self, there is no delusion, no realization, no Buddha, no sentient beings, no birth and death. We do not need to worry about exactly what is going on here—though one needs to understand this if one is to understand why Dōgen might be expressing these views. Even without this, it is quite clear that he is endorsing a contradiction.27 Does he do this with the contradiction involved in talking of the ineffable? 2.5 Katto To find an answer to this question, we can go to the Shōbōgenzō lecture entitled Katto (Kudzo and Wisteria, 葛藤). Note that Kudzo and Wisteria are both plants that grow vines, which climb by wrapping themselves round other things. Back to this in a moment. In Zen thought there is a familiar story about how Bodhidharma elected his successor, Huike. In Katto, Dōgen relates this as follows:28 The twenty-eighth patriarch said to his disciples, “As the time is drawing near [for me to transmit the Dharma to my successor], please tell me how you express it.” Daofu responded first, “According to my current understanding, we should neither cling to words and letters, nor abandon them altogether, but use them as instruments of the Dao [CP: way].” The master responded, “You express my skin.” Then the nun Zongzhi, said, “As I now see it [the Dharma] is like Ānanda’s viewing the Buddha-land of Akshobhya, seeing it once and never seeing it again.” The master responded, “You express my flesh.” Daoyou said, “The four elements are emptiness, and the five skandhas are non-being. But in my view, there is not a single dharma to be expressed.” The master said, “You express my bones.” Finally, Huike prostrated himself three times and stood [silently]. The master said, “You express my marrow.”

288  Filippo Casati and Graham Priest The standard interpretation of the story is that each of the respondents says something acceptable, each getting closer to the essence of things. By his silence, Huike indicates the ultimate, and so ineffable. On the basis of this, Bodhidharma makes him his successor. Dōgen’s interpretation of the story is, however, quite different. He says (ibid.): You should realize that the first patriarch’s expression, “skin, flesh, bones, marrow,” does not refer to the superficiality of depth [or understanding]. Although there may remain a [provisional] distinction between superior and inferior understanding, [each of the four disciples] expressed the first patriarch in his entirety. When Bodhidharma says “you express my marrow” or “you express my bones,” he is using various pedagogical devices that are pertinent to particular people, or methods of instruction that may or may not be apply to different levels of understanding. It is the same as Śākyamuni’s holding up the udambarra flower [to Mahākāśyapa],29 or the transmission of the sacred robe [symbolic of the transmission of enlightenment]. What Bodhidharma said to the four disciples is fundamentally the selfsame expression. Although it is fundamentally the selfsame expression, since there are necessarily four ways of understanding it, he did not express it in one way alone. But even though each of the four ways of understanding is partial or one-sided, the way of the patriarchs ever remains the way of the patriarchs. Dōgen’s point is that the replies of each of the disciples do not indicate differences of depth, but merely different ways of saying the same thing (and each might be appropriate on different occasions). The speech of the first three disciples and the silence of Huike, then, are all equivalent. So silence is not privileged over speech. After all, a body is an integrated whole, and each part of it is necessary for the others. The replies of the disciples are mutually co-dependent, like the parts of the body. So what has this to do with kudzo and wisteria? The discussion of the transmission takes off from the following passage (ibid.): My late master [Ruijing] [CP: 如淨] once said: “The vine of a gourd coils around the vine of a[nother] gourd like a wisteria vine.” I have never heard this saying from anyone else of the past or the present. The first time I heard this was from my late master. When he said, “the vine of a gourd coils round the vine of a[nother] gourd,” this refers to studying the Buddhas and patriarchs directly from the Buddhas and patriarchs, and to the transmission of the Buddhas and patriarchs directly to the Buddhas and patriarchs. That is, it refers to the direct transmission from mind-to-mind.

Heidegger and Dō gen on the Ineffable  289 The direct transmission is the silence of Hiuke. That is the vine of one gourd. Language is the vine of the other. The being of each of these is necessary for the being of the other. And both, in their own way, say the same thing. We have here a reprise of the Vimalakīrti Nirdeśa Sūtra. It is exactly the point that the goddess is making when she says that words are not internal, not external, and not anywhere else. And it is exactly the final entanglement between Mañjuśrī and Vimalakīrti. Dōgen is saying that one can talk of the ineffable; indeed, the ineffable and the effable each requires the other.

3. Heidegger 3.1 Being Having seen Dōgen’s approach to the problem of ineffability, let us now turn to Heidegger.30 As we will see in this section, even though these two philosophers come from very different traditions, they share many ideas. First, both Heidegger and Dōgen agree that something is ineffable. For Dōgen, this is ultimate reality; for Heidegger, it is being. Secondly, they both believe that ineffability leads to a contradiction. Thirdly, they both endorse dialetheism concerning ineffability: they think that the contradiction implied by talking about the ineffable is true. And last, they both take the speech of effability and the silence of ineffability to be entangled. Martin Heidegger (1889–1976) is clearly one of the most influential philosophers of the 20th century. Hans-Georg Gadamer, one of the most famous of Heidegger’s students, describes his teacher as the last shaman, owing to his ability to enchant the students. His classes were always full and the course that he gave at the University of Freiburg, during the summer semester of 1935, did not constitute an exception. On that occasion, Heidegger began his Introduction to Metaphysics with the following question: “Why are there entities at all instead of nothing?”31 Of course, it is a trivial fact that we are constantly surrounded by entities, such as books, dreams, mathematical theorems, and cherry-trees. What is not trivial is to understand why there are all these entities and what makes them be. An easy answer could be that, as the redness of the rose makes the rose red, the being of all entities makes all entities be. But, then, what is the being of all entities? This is the well-known Seinsfrage, the question of being; and from the beginning of his philosophical work, Heidegger aimed to answer it. In Being and Time, he writes:32 what is asked about is being, . . . that on the basis of which entities are already understood, however we may discuss them in detail. So, what is the being of all entities? According to Heidegger, being has two main features. The first one is that being is the ‘being an entity’ of all

290  Filippo Casati and Graham Priest entities—the Seiendsein. As such, it “determines entities as entities.”33 Being is that in virtue of which all entities are something and not nothing. Since Heidegger interprets the relation between being and entities as a grounding relation, then being can be also characterized as the ground of all entities—the Grund. He writes: “being is intrinsically ground-like—what gives ground” to all entities.34 The second feature of being is that being itself is not an entity. According to Heidegger, there is an ontological difference between being and what being makes be, namely all entities. In Being and Time, we can read:35 The being of entities ‘is’ not itself an entity. If we are to understand the problem of being, our first philosophical step consists in . . . not ‘telling a story’, that is to say, in not defining entities as entities by tracing them back in their origins to some other entities—as if being had the character of a possible entity. Even though Heidegger does not give any reason to endorse the ontological difference, it is possible to defend this position appealing to both a metaphysical argument and a grammatical argument. The metaphysical argument aims to show that, without an ontological difference between being and entities, a vicious infinite regress is generated. To see this, consider the case in which the ontological difference does not hold, namely the case in which everything, including being, is an entity. As we have noted before, being grounds all entities; and all entities, in order to be something and not nothing, need to be grounded in something else. If so, since everything is an entity, including being, then being needs to be grounded in something else as well. Let’s say, then, that being is grounded in being2. However, as everything is an entity, being2 must be an entity as well and, as such, it must be grounded in something else too. Let’s say that being2 is grounded in being3. At this point, it is easy to see that we are off on an infinite regress. Such an infinite regress is vicious if one believes that, in our explanans, we cannot invoke that very thing for which we are seeking an explanation. Indeed, in explaining why there are entities in the first place, we are always invoking an entity (namely being, being2, being3. . . .). Such a vicious infinite regress is broken if we assume that being, which grounds all entities, is not itself an entity.36 Let’s move on with the second argument—the grammatical one.37 Consider a proposition such as ‘the wall is’. The noun ‘wall’ refers to an entity, namely the wall behind me. But what shall we say about ‘is’? If we assume that ‘is’ (namely ‘being’) refers to an entity, then the proposition ‘the wall is’ would be nothing more than a simple list of two entities: the wall and is (namely being). However, this cannot be the case because, in an obvious sense, the proposition ‘the well is’ has a meaning that a list of two entities (such as the wall and is) does not have. Therefore, it is possible to conclude that the ‘is’ of the proposition ‘the wall is’ does not refer to any entity

Heidegger and Dō gen on the Ineffable  291 because being is not an entity. As Heidegger himself suggests in his The Principle of Reason:38 only an entity ‘is’; the ‘is’ itself—being—‘is’ not. The wall in front of you and behind me is. It immediately shows itself to us as something present. But where is its ‘is’? Where should we seek the presencing of the wall? Probably these questions always run awry. In the metaphysical picture defended by Heidegger, the influences of neoplatonism and medieval philosophy are clear. On the one hand, as for Plotinus’s One, being is that in virtue of which all entities are. Being is the ground of all entities. On the other hand, as for Angelus Silesius’s God and Meister Eckhart’s Gottheit, being is completely transcendental. It is beyond all entities because being is not itself an entity. In other words, Heidegger thinks that being is that in virtue of which the world is, and the world needs to be interpreted as the collection of all entities. Moreover, since being is not an entity, being transcends the world. Therefore, being (namely the reason in virtue of which everything is, including the world itself) is not part of the world.39 3.2 The Paradox of Being As presented till now, the path taken by Heidegger to answer the question of being does not seem particularly troubled. To understand why this is not the case, it is necessary to focus our attention on what, according to Heidegger, should be taken as an entity. In Being and Time, we read that an entity is:40 everything we can talk about, everything we have in view, everything towards which we comport ourselves in any way. If we speak, dream, or fear about something, we speak, dream or fear about some thing, namely an entity.41 Since we say that “ ‘the earth is’, ‘the lecture is in the auditorium’, ‘This man is from Swabia’, ‘the cup is of silver’,”42 the earth, the lecture in the auditorium, the man from Swabia, and that cup made out of silver are all entities. Indeed43 when we say something ‘is’ and ‘is such and so’, then that something is, in such an utterance, represented as an entity. Now, assuming this definition of entity, the problem that Heidegger faces is evident. According to the ontological difference, being is what determines entities as entities, without being an entity. Nevertheless, in saying that being determines entities as entities, we treat being as an entity because we talk about it and, above all, we describe it as ‘such and so’. Therefore, being

292  Filippo Casati and Graham Priest is not an entity (because of the ontological difference) and being is an entity (because we talk about it and we describe it as ‘such and so’). Bad news never comes alone. Indeed, it is not only the case that we cannot answer the question of being because it is impossible to think and speak about it. We cannot even ask the question of being. According to Heidegger, any question presupposes an entity the question is about. For instance, if we ask something about a city, then we ask something about a thing, an entity. It follows that, since questions are always questions about some things (namely entities) and since being is not an entity, nothing can be asked about being. Using Heidegger’s words in Being and Time:44 Inquiry, as a kind of seeking, must be guided beforehand by what is sought. So the meaning of Being must already be available to us in some way. As we have intimated, we always conduct our activities in an understanding of Being. Out of this understanding arise both the explicit question of the meaning of Being and the tendency that leads us towards its conception. We do not know what ‘Being’ means. But even if we ask, ‘What is ‘Being’?’, we keep within an understanding of the ‘is’, though we are unable to fix conceptionally what that ‘is’ signifies. So, Heidegger finds himself stuck in a paradoxical situation. On the one hand, being is ineffable. Indeed, if we try to speak about it, we talk about some thing (namely an entity) and being is not a thing (an entity). On the other hand, being is not ineffable because we can and we do talk about it. Moreover, since everything we talk about is an entity, being is an entity as well. Therefore, being is ineffable (because it is not an entity) and being is not ineffable (because it is an entity).45 Heidegger is perfectly aware of this situation, and that such a contradiction leads his whole metaphysical project to a dead-end. Indeed, the aim of Heidegger’s metaphysics is to answer the question of being; and in order to answer the question of being, it is necessary to talk and think about being. However, as we have seen, talking and thinking about being leads one to claim something contradictory. This makes any attempt of answering the question of being meaningless because:46 a contradictory speech is an offense against the fundamental rule of speech (logos), against logic. . . . Logic is taken as a tribunal, secure for all eternity, and it goes without saying that no rational human being will call into doubt its authority as the first and last court of appeal. Whoever speaks against logic is suspected, explicitly or implicitly, of arbitrariness. It is not easy to understand what Heidegger has in mind when he talks about logic. However, for the purpose of the present essay, it is enough to say that, according to Heidegger, logic is “a set of rules” for “a good way of

Heidegger and Dō gen on the Ineffable  293 reasoning” grounded on “two main principles: the principle of non-contradiction and the principle of identity.”47 As we have seen, answering the question of being violates one of the fundamental principles of good reasoning, namely the principle of non-contradiction. Since Heidegger says here that any violation of such a principle cannot be accepted by any rational human being, then his own theory cannot be accepted by any rational human being either. He writes: “[CP: Talking and thinking about being] is contradictory and, therefore, senseless.”48 Since Heidegger does not want to abandon any of his metaphysical assumptions and since he endorses the account of logic sketched above, he coherently concludes that thinking and speaking about being is impossible. Being remains unfathomable. Tragically, Heidegger faces the evidence that his whole metaphysical project is self-defeating and, until the well-known methodological turn of his thought, he did not know how to escape from such a dead-end. Such a turn is called the Kehre and it took place around the 1930s. 3.3 Beyng As the problem has been framed here, Heidegger has two available options. Either he revises some of his metaphysical assumptions or he revises his logical assumption. After the Kehre, Heidegger started to focus his attention on the first option, trying to find a way to talk and think about being without turning it into an entity. According to Heidegger, such a way is represented by poetry and, more generally, art, because a genuine work of art can show being without saying anything about it. The work of art opens up in its own way the being of entities. This opening up, i.e., this revealing, i.e., the truth of entities, happens in the work. In the art work, the truth of entities has set itself to work. Art is truth setting itself to work.49 A solution to the question of being may, therefore, lie in how art can open people’s eyes in this way. At this point two remarks are necessary. First, it is important to state clearly that it is unquestionable that Heidegger tried to solve the problem of the ineffability of being by endorsing a poetic way of talking and thinking about it. As has been extensively discussed in the secondary literature, Heidegger thought that poetic language was a way of referring to being that did not imply any reification or objectification of being itself.50 We do not wish to deny that Heidegger pursued this strategy. However, the fact that Heidegger engaged with a poetic solution to the ineffability of being does not mean that Heidegger engaged with only this solution. The second remark is about the efficacy of Heidegger’s poetic solution. A discerning eye will itself still perceive a problem here. Never mind

294  Filippo Casati and Graham Priest answering the question of being; as already noted, if one cannot refer to being with a noun phrase, one cannot even ask it. Indeed, for exactly this reason, one can say nothing at all of being. Yet, Heidegger’s own works are replete with statements about being. To bend a comment from Russell’s introduction to Wittgenstein’s Tractatus—which finds itself with a similar aporia:51 Everything involved in talking of being cannot, grammatically, be said. What may give some hesitation about this fact is that, despite his arguments to the contrary, Mr. Heidegger manages to say a good deal about what cannot be said. Nor, we note, is this necessarily incompatible with poetry/art showing us being; for, pace Wittgenstein, what can be shown can often be said. So we are left with the second option: challenging the receive logic. Did Heidegger ever consider this? Did he ever try to abandon the Principle of Non-Contradiction? According to the secondary literature, the answer is negative. In what follows, we disagree with this interpretation, showing that, after the Kehre, Heidegger challenges the Principle of Non-Contradiction and openly endorses dialetheism (the view according to which some contradictions are true). The complete and clear realization that being requires us to abandon the Principle of Non-Contradiction was formulated in years of both private and public philosophical attempts. To begin with, Heidegger starts to cast some doubts about the Principle of Non-Contradiction in his lecture of 1929, What Is Metaphysics? And he asks: “Are we allowed to tamper with the rules of logic?.”52 Nevertheless, the first essay in which Heidegger seems to endorse a dialetheic approach to the paradox of being, accepting its contradictory nature, is contained in his Introduction to Metaphysics. He writes:53 the word ‘being’ is thus indefinite in its meaning, and nevertheless we understand it definitely. ‘Being’ proves to be extremely definite and completely indefinite. According to the usual logic, we are here on obvious contradiction. But something contradictory cannot be. There is no square circle. And yet, there is this contradiction: being as definite and completely indefinite. We see, if we do not deceive ourselves, and if for a moment amid all the day’s hustle and bustle we have time to see, that we are standing in the midst of this contradiction. This standing of ours is more actual than just about anything else that we call actual—more actual than dogs and cats, automobiles and newspapers. In this paragraph, Heidegger rephrases the paradox of being. On the one hand, he claims that the word ‘being’ refers to something that does not have any determination (something about which nothing can be either said or thought because there are no determinations to be said and thought about it). On the other hand, he claims that the word ‘being’ refers to

Heidegger and Dō gen on the Ineffable  295 something that has some determinations (something about which it can be said or thought: at least, the determination of not having any determinations). In other words, being is indeterminate (it has no determination at all) and it is determinate (it has the determination of not having any determination at all). It is important to notice that, in this paragraph, Heidegger does not simply rephrase the contradiction of being. He also states that this contradiction is actual. Here, Heidegger explicitly endorses the idea that the contradiction of being has to be accepted as true because unavoidable. According to Heidegger, such a contradiction is as actual and real as all the other actual and real things in the world. It is as actual and real as dogs and cats, automobiles and newspapers. It would be intuitive to expect that, as he explicitly accepts the idea that there is one actual contradiction (namely the contradiction of being), Heidegger systematically accepts the idea that contradictions are not necessarily unacceptable too. Unfortunately, this is not the case. The idea that the Principle of Non-Contradiction should (or simply could) be abandoned, accepting the contradiction of being as true, is not consistently presented throughout his Introduction to Metaphysics. Besides the paragraph just noted, there are no other significant indications of this direction of thought. However, some years after the publication of Introduction to Metaphysics, the dialetheic solution to the paradox of being was systematically presented in his Contributions to Philosophy—a philosophical diary written between 1937 and 1938, but published only after Heidegger’s death. In this work, Heidegger presents a full defence of the position according to which being should be taken to be both an entity and not an entity. In order to mark the difference between his old consistent account of being and the new inconsistent one, he starts to write being [Sein] as beyng [Seyn]. Following Heidegger, in discussing his dialetheic approach to the paradox of being, we will start writing being as beyng, too. In the Contributions to Philosophy, Heidegger claims that metaphysics needs a new beginning: These ‘contributions’ question along a way which is first paved by the transition to the another beginning, the one Western thought is now entering. . . . [CP:An]other beginning must be attempted.54 This new beginning is represented by the introduction of the ‘event’ [Ereignis] which is, the self-eliciting and self-mediating center in which all essential occurrence of the truth of beyng must be thought back in advance.55 The event is described as the occasion in which the truth of beyng is disclosed, and such a truth is disclosed though the human being’s “thinking

296  Filippo Casati and Graham Priest of beyng.”56 Here, the truth of beyng needs to be interpreted as something true about beyng. In the event, the human being, thinking about beyng, reveals something true about it. For this reason, Heidegger metaphorically characterized the event as the human being’s appropriation (Er-eignung) of the truth of beyng. At this point, even though the contradictory nature of beyng has not been explicitly accepted yet, a dialetheic solution is definitely implied by Heidegger’s event. On the one hand, according to the metaphysical premises of Heidegger, thinking and speaking about beyng leads to a contradiction. Such metaphysical premises are certainly not rejected in Contributions to Philosophy. On the other hand, in the event, thinking about beyng discloses something true about it. It seems to follow that the truth of beyng precisely consists in thinking something contradictory, but still true, about beyng itself. This position becomes explicit when the real content of the event is properly described. According to Heidegger, in the event, not only is beyng held as the complete opposite of an entity (because, according to the ontological difference, beyng is not an entity), but beyng is also held as something that is not the complete opposite of an entity (because, since we think and speak about it, beyng is an entity too). This is the reason why, even though the ontological difference still holds, ensuring the fact that beyng is not an entity, Heidegger thinks that, “beings [CP: or entities] are, Beyng essentially occurs” as well.57 If so, according to the truth of beyng disclosed in the event, beyng itself is both an entity and not. Consistently with this position, in Contributions to Philosophy, paragraph number 47, Heidegger questions the idea that it is necessary to choose between the following two options: either beyng is an entity or beyng is not an entity. Heidegger wants to question exactly this either/or. He wants to challenge the necessity of choosing one of these two options because choosing both would mean to claim something contradictory and, thus, senseless. He provocatively asks:58 Whence the either-or? Whence the only this or only that? Whence the unavoidability of this way or else that way? Is there not still a third (. . .)? One paragraph later, Heidegger answers that the third available option is to challenge the Principle of Non-Contradiction, accepting that beyng is both an entity and not. For this reason, Heidegger writes that the truth of beyng is represented by “the being of a nonbeings [CP: non-entities].”59 Consistently with what we have said before, this truth is contradictory because a non-entity, namely something that is not an entity (like beyng), cannot be. As such, there cannot be any being of a non-entity either. Nevertheless, in Heidegger’s event, the essence of beyng reveals being itself as something that is not (because, according to the ontological difference, it is a non-entity)

Heidegger and Dō gen on the Ineffable  297 but, nonetheless, is. Therefore, beyng is (an entity) and not. Such a contradictory truth about beyng is fully endorsed by Heidegger: Non-being as a mode of being: it is and yet is not. And likewise being: permeated with the ‘not’ and yet it is.60 Those who fancy themselves only too clever and immediately uncover a contradiction here, since indeed nonbeings [CP: such as beyng] cannot ‘be’, are thinking in much too narrow way with their ‘non-contradiction’ as the measure of [CP: beyng, namely] the essence of beings.61 According to the late Heidegger, then, thinking the essence of entities requires us to revise our logical beliefs, abandoning the Principle of NonContradiction. So, there are books, dreams, mathematical theorems, and cherry-trees because of beyng. Beyng makes all entities be, while beyng itself is an entity and not. Heidegger’s rejection of the Principle of Non-Contradiction is at its most explicit in the Contributions to Philosophy. However, there are certainly many other allusions to the matter in his later work. Thus, in a seminar given at the University of Freiburg during the summer semester 1934, Heidegger claims that his philosophy has “the necessary task of a shaking up of logic.”62 About six years later, in a collection of essays entitled The History of Beyng, written between 1938 and 1940, we find:63 a contradiction is not a refutation. . ., but rather fathoming the ground of an inceptual fundamental position within the truth of beyng. And again, in an essay entitled What Is a Thing? published in 1962, Heidegger writes that:64 the Principle of Non-Contradiction is not a basic principle of metaphysics. . . . Logic cannot be the fundamental science for metaphysics. The statements are, perhaps, somewhat coy, compared with the remarks on the matter in the Contributions to Philosophy; but given these remarks, their intent is clear. 3.4 Silence So far so good; but it still remains the case that beyng is ineffable; and if it is ineffable, approaches to it must deploy silence. The matter is not lost on Heidegger. In his Contributions to Philosophy, Heidegger reaches the following conclusion: given the right conditions, keeping silent about being helps us to get closer to the truth of being without completely reaching it. This is also why Heidegger claims that “the other beginning is carried out as bearing silence.”65 As we explained in the previous section, the new

298  Filippo Casati and Graham Priest beginning is the one in which, through the event, the truth of beyng is disclosed. Of course, bearing silence cannot fully represent what happens in the new beginning because, exactly in the new beginning, Heidegger is not silent at all. On the contrary, he speaks in great length about beyng and its truth. So, how is the new beginning related to silence? What is the connection between the action of bearing silence about being and the truth of being expressed in the new beginning of metaphysics, celebrated by Heidegger himself? According to Heidegger, in some specific circumstances, silence communicates something or, in other words, it has a content. Silence is not silent at all. This idea is already endorsed by Heidegger before the Kehre. In Being and Time, he writes:66 Keeping silent is another essential possibility of discourse, and it has the same existential foundation. In talking with one another, the person who keeps silent can ‘make one understand’ (that is, they can develop an understanding), and he can do so more authentically than the person who is never short of words. From this quotation, it is clear that, according to Heidegger, a silent person can still convey something. Clearly, however, not all kinds of silence are equally communicative. One person might be silent because they are sleeping, and another might be silent because they do not understand the topic of a conversation. In these cases, the silence does not communicate much. In his Contributions to Philosophy, Heidegger seems to suggest that, in order to move closer to the truth of being, silence and language need to be necessarily entangled: on the one hand, silence needs language and, on the other hand, language needs silence. Let’s begin discussing why silence needs language. According to Heidegger, in order to be able to convey something about being, the silence requires us to speak about being. Keeping silent about being makes sense if and only if we speak and ask about being in the first place. In Heidegger’s words:67 When this restraint [CP: the restraint of language into silence] reaches words, what is said is always the event [CP: namely the truth of being]. . . . The saying that bears silence is what grounds [CP: the event or the truth of being]. The silence that is able to communicate something about being is one that “grows only out of restraint [CP: namely the restraint of talking about being].”68 As such, it needs language that talks about being in the first place. Exactly the context in which the silence is placed, namely the context of a discussion about being, gives the right meaning to the silence that follows such a discussion.

Heidegger and Dō gen on the Ineffable  299 Now the other direction: according to Heidegger, language needs silence as well. This is the case because it is the silence that shows the failure of the language in talking about being. In other words, the failure of language is completely revealed by the necessary silence to which a speaker is forced when they understand that being is actually ineffable and, as such, nothing can be expressed about it. Moreover, such a failure is important because, according to Heidegger, it discloses something about being itself, namely the impossibility of talking about it and the relative limits of the language. Heidegger expresses this idea in the following way:69 Words fail us; they do so originally and not merely occasionally, whereby some discourse or assertion could indeed be carried out but is left unuttered. . . . Words do not yet come to speech at all, but it is precisely in failing us that they arrive at the first leap. This failing is the event as intimation or incursion of beyng. To conclude, Heidegger holds that silence and language are necessary entangled. This is also the reason why Heidegger writes that the truth of being is revealed in the union or mutual entanglement of “bearing silence and questioning [CP: being].”70 In order to get closer to the truth of being, the pair of speaking and keeping silent is necessary: each requires the other.

4.  And So to Nothingness What we have seen so far is the following. Both Dōgen and Heidegger think that there is something that is ineffable; both talk about it; both recognise the contradictory nature of this, and accept the contradiction. Moreover, both see that the speech of ineffability and the silence of ineffability are entangled: each necessitates the other. Both Heidegger and Dōgen are philosophers who are sometimes difficult to understand; and the parts of their thought about which we have been writing are certainly so.71 Seeing the thought of each through the lens of the other can help us to understand both. It might seem, none the less, that, for all the similarities, Dōgen and Heidegger are dealing with quite different topics. Heidegger’s being, after all, would appear to be quite different from Zen’s ultimate reality. The appearance is deceptive. In the thought of both, what they are concerned with can be seen as exactly the same thing: nothingness. In this final section, we shall see why. 4.1  Heidegger: Being and Nothingness Let us start, this time, with Heidegger. For him the connection between being and nothingness is clear: he states that they are identical. In What Is Metaphysics? Heidegger states the matter as follows:72

300  Filippo Casati and Graham Priest ‘Pure Being and pure Nothing are therefore the same.’ This proposition of Hegel’s (Science of Logic, vol. I, Werke III, 74) is correct. Being and the nothing do belong together, not because both—from the point of view of the Hegelian concept of thought—agree in their indeterminateness and immediacy, but rather because Being itself is essentially finite and reveals itself only in the transcendence of Dasein which is held out into the nothing. Heidegger’s reason for supposing that being and nothing are identical is somewhat opaque, but appears to be the following simple argument: Being is what it is that makes beings be. Nothing is what it is that makes beings be. Hence, being is nothing. The first premise is true by definition. The conclusion follows validly if there is just one thing that makes beings be. So perhaps the most contentious part of the argument is the second premise. The thought here is that what makes something be is the fact that that it “stands out against nothingness.” So if there were no nothingness, there could be no beings either. As Heidegger puts it:73 In the clear night of the nothing of anxiety the original openness of beings as such arises: they are beings—and not nothing. But this ‘and not nothing’ we add in our talk is not some kind of appended clarification. Rather it makes possible in advance the revelation of beings in general. The essence of the originally nihilating nothing lies in this, that it brings Dasein for the first time before beings as such. Or again:74 The nothing is neither an object nor any being at all. The nothing comes forward neither for itself nor next to beings, to which it would, as it were, adhere. For human existence the nothing makes possible the openedness of beings as such. The nothing does not merely serve as the counterconcept of beings; rather it originally belongs to their essential unfoldings as such. In the Being of beings the nihilation of the nothing occurs. This is not the place to discuss the details of these views. Here, we merely note Heidegger’s conclusion: that being and nothing are identical. 4.2  Heidegger: Ineffability and Nothingness If being is nothingness, and being is ineffable, then so too, presumably, is nothingness. But, in fact, there are independent reasons for supposing that nothingness (and so being) is ineffable.

Heidegger and Dō gen on the Ineffable  301 To say something of something, it has to be an object—a thing of which one can predicate some characterisation. But this is exactly what nothingness is not. By definition, nothingness is the absence of all objects. There is then no thing there of which to predicate anything. As Heidegger puts it:75 What is the nothing? Our very first approach to the question has something unusual about it. In our asking we posit the nothing in advance as something that ‘is’ such and such; we posit it as a being. But that is exactly what it is distinguished from. Interrogating the nothing—­asking what, and how it, the nothing, is—turns what is interrogated into its opposite. The question deprives itself of its own object. Accordingly, every answer to the question is also impossible from the start. For it necessarily assumes the form: the nothing ‘is’ this or that. With regard to the nothing question and answer are alike inherently absurd. Nothingness, then, is ineffable. Of course, it is just as obvious that one can say something about nothingness. We and Heidegger just have. We have been over this matter already, however. The point of this section is simply to establish that, for Heidegger, being is the same as nothingness, which we have now done. 4.3  Nishitani: Absolute Nothingness Let us, then, return to Dōgen. According to Heidegger, nothingness is the absence of all things:76 If beings are taken in the sense of objects and objectively present things . . . nothingness signifies the utter negation of beings understood in this sense. Such is also the ineffable ultimate reality of Mahāyāna Buddhism. For, as we noted in 4.3, this ultimate transcends all dualities; and if it contained objects, it would contain dualities (thises and thats). The nature of nothingness comes out most clearly in the East Asian forms of Buddhism, in which Dōgen was working, and which was heavily influenced by Daoism.77 In Daoism (or, at least, the version of it which influenced Buddhism), behind the myriad things of the phenomenal world there is an originating principle, dao 道. This can be no particular thing, or it could not engendeer all things. As Wang Bi (226–249 CE, an influential Neo-Daoist) puts it:78 The way things come into existence and efficacy [gong] comes about is that things arise from the formless [wuxing] and that efficacy emanates from the nameless [wuming]. The formless and the nameless [the Dao]

302  Filippo Casati and Graham Priest is the progenitor of the myriad things. It is neither warm nor cool and makes neither the note gong nor the note shang. . . . If it were warm, it could not be cold; if it were the note gong, it could not be the note shang. If it had a form, it would necessarily possess the means of being distinguished from other things; if it made a sound, it would necessarily belong among other sounds. Unlike the beings (Chin: you, 有) comprising the myriad things, then, it is nothingness, no thing (Chin: wu; Jap: mu, 無). The ineffability of nothingness, and the contradiction involved in speaking of it, comes out very clearly in the thought of Nishitani Keiji (西谷 啓治, 1900–1990). Nishitani was professor of philosophy and religion at Kyoto University, and one of the most important members of the Kyoto School of Philosophy. Members of this school drew on aspects of both Zen philosophy and Western thought. Nishitani’s thought is heavily influenced by both Dōgen and Heidegger, with whom he studied in the late 1930s.79 Nishitani often refers to the nothingness of ultimate reality as absolute nothingness (zettai mu, 絶対無) to distinguish it from the nothingness of simply non-being (nihility). Absolute nothingness is “beyond being and non-being.” What is and what is not, constitute conventional reality; absolute nothingness, being ultimate reality, transcends both. Nishitani puts the point this way:80 Viewed in terms of this process [CP: of moving from being to absolute nothingness], śūnyata81 represents the endpoint of an orientation to negation. It can be termed absolute negativity, inasmuch as it is a standpoint that has negated and thereby transcended nihility, which was itself the transcendence-through-negation of all being. It can also be termed as an absolute transcendence of being, as it absolutely denies and distances itself from any standpoint shackled in any way whatever to being.

4.4  Nishitani: Ineffability and Nothingness Unsurprisingly, then, and as the last quotation itself suggests, Nishitani endorses the ineffability of this nothingness. As he says:82 On the field of emptiness, however, the selflessness of a thing cannot be expressed simply in terms of “being one thing or another.” It is rather something laid bare as something that cannot on the whole be expressed in the ordinary language of reason, nor for that matter in any language containing logical form.

Heidegger and Dō gen on the Ineffable  303 There is of course, no language whose sentences do not have logical form— except perhaps simply pointing (tathā), if that be a language. Nishitani is well aware, however, that he has been talking about nothingness in a language (Japanese—of course, a language whose sentences have logical form). He embraces the paradox. The quotation continues:83 Should we be forced to put it into words all the same, we can only express it in terms of paradox, such as: “It is not this thing or that, therefore it is this thing or that.” Nishitani’s thought is frequently dialetheic. Thus, for example, we have:84 In other words, true nirvāna appears as samsāra-sive-nirvāna. Here life is sheer life and yet thoroughly paradoxical. We can speak, for example, of essentiality in its true essence as non-essentiality. If we could not speak in such terms as these, life would not truly be life. It would not be life at once truly eternal and truly temporal. Indeed, a few lines later Nishitani quotes Dōgen with approval on the matter:85 In Dōgen’s words, “Birth itself is non-birth; extinction itself is non-extinction.” No reference to Dōgen is given, but we take this to be a reference to the passage from Shoji which we quoted in 2.4. Nishitani, just as much as Dōgen, then, deploys contradiction; and deploys it, in particular, to endorse speaking of the ineffable. The main point of this section is, however, simply to show that for these thinkers ultimate reality is nothingness. For both Heidegger and the Zen philosophers, then, the topic of the aporia which has been the central concern of our paper may be seen as exactly the same: nothingness.

5. Conclusion Heidegger and Dōgen might seem to be very different philosophers, with different concerns and agendas. And so they are. Each is concerned with many things which do not concern the other. However, at the heart of each of their projects is a core notion: being for Heidegger, and ultimate reality for Dōgen; and in their thought, as we saw with the help of Nishitani, both of these, under inspection, morph into nothingness. Moreover, for both of them this is ineffable; and so both of them face the issue of talking about the ineffable. Fainter hearts—at least, hearts in thrall to Aristotle—might have

304  Filippo Casati and Graham Priest tried to wriggle out of the contradiction. Both, however, have the courage of their arguments and endorse it. One can speak about certain ineffable notions. Moreover, both go beyond simply endorsing the contradiction involved. Both grapple with what this means for the relationship between speech and silence in the matter; and both come to the conclusion that the two are indelibly entangled. What one might not have expected, but what we have now seen, is that the thoughts of Heidegger and Dōgen are themselves indelibly entangled.

Notes 1 Miller (1969, pp. 81, 82). 2 E.g., Fine (1975), Tye (1994). 3 The qualification is required. It is natural to suppose that the number three is neither red nor not red; but this is not a case of indeterminacy, just of a category mis-match. 4 For an informative overview of the relation between Heidegger and the Zen tradition, see May (2005). 5 See Priest (2002, esp., p. 227). 6 On Indian Buddhism, see Siderits (2007). On Mahāyāna Buddhism (Indian and East Asian), see Williams (2009). 7 Note that translators of Buddhist texts usually translate the Sanskrit word satya as truth; however, it can also be apt to translate it as reality, which is much more sensible in the present context. 8 MMK, XVIII: 9. Garfield (1995, p. 49). 9 Garfield (1995, p. 2). 10 MMK XIII:8. Garfield (1995, p. 36). 11 MMK XXII: 11, 12. Garfield (1995, p. 61 f.). 12 Conze (1973, p. 12). The Prajñāpāramitā Sūtras were a class of sūtras which appeared around the turn of the Common Era, and heralded the advent of Mahāyāna Buddhism. 13 Though not, interestingly, Nāgārjuna himself. The MMK does not comment on the matter. 14 Gorampa Sonam Senge (1429–1489) is one of the central philosophers in the Sakya school of Tibetan Buddhism. 15 The translation is taken from Kassor (2013, p. 401). The ‘four extremes’ is a reference to the trope of Buddhist logic called the catuṣkoṭi, which enumerates the four kinds of things that can be said about something. 16 Kassor (2013, p. 406). She notes (in correspondence) that the nomenclature for the distinction employed here is not Gorampa’s, but is that of another thinker. He himself calls the distinction one between the ultimate that is realized, and the ultimate that is taught. 17 Thurman (2014, p. 59). 18 Ibid., p. 77. 19 See, further, Garfield (2002). 20 Such as the study of kōan—certain verbal puzzles—and shock tactics on the part of the teacher, such as shouting or striking. 21 On Chan in general, see Hershock (2015). On Dōgen in particular, see Abe (1992). 22 Tendai is one of the other major schools of Chinese Buddhism. 23 See the introduction to Cleary (1986, esp. pp. 5 ff.). 24 Cleary (1986, p. 39). The italics are Cleary’s. In what follows, translators’ interpolations are marked with square brackets. Ours are marked with initials [CP: thus].

Heidegger and Dō gen on the Ineffable  305 5 Tanahashi (1985, p. 74). 2 26 Ibid., p. 69. 27 This translation and interpretation of the passages is defended in Deguchi, Garfield, and Priest (2013). 28 Heine (2009, pp. 151–152). In this context, ‘Dharma’ means something like true teaching. 29 This is an allusion to a Chan story about how the Buddha himself elected his successor. Whilst with a bunch of his disciples, he simply held up a flower. None of the disciples knew what to do, except Mahākāśyapa, who just smiled. He had grasped the ineffable, which the Buddha was conveying; the Buddha elected him. ‘Śākyamuni’ is a name for the Buddha, and means sage of the Śākyas, the Buddha’s clan. 30 Much of the material in this section comes from Casati (2016). 31 Fried and Polt (2000, p. 1). 32 Macquarrie and Robinson (1962, pp. 25–26). 33 Ibid., p. 25. 34 Stambaugh (1985, p. 170). 35 Macquarrie and Robinson (1962, p. 26). 36 A similar argument in favour of the ontological difference is presented by Nicholson (1996). 37 This can be found in Priest (2002, 15.3). 38 Lilly (1991, p. 51). 39 Concerning the relation between Heidegger and neo-platonism, see his lectures on Greek philosophy (Rojcewicz (2007)) and Plato (Rojcewicz and Schuwer (2003)). See, also, Cimino (2005) and Narbonne (2001). Concerning Heidegger and medieval philosophy, see Fritsch and Gosetti-Ferencei (2010) and Caputo (1986). 40 Macquarrie and Robinson (1962, p. 26). 41 Heidegger uses different terms to talk about entities [Seiendes]: for instance, he uses the term ‘thing’ [Ding], the term ‘being’, and the term ‘object’ [Objectum]. All these terms have different (phenomenological) meanings for him. Nevertheless, this is not relevant here, and for our purposes we may simply take these terms as synonyms. 42 Fried and Polt (2000, p. 93). 43 Lilly (1991, p. 51). 44 Macquarrie and Robinson (1962, p. 25). 45 The paradox faced by Heidegger has been extensively discussed in the secondary literature. For instance, see Casati and Wheeler (2016), Witherspoon (2002), Moore (2012), Priest (2002). 46 Fried and Polt (2000, pp. 25–27). 47 Gregory and Unna (2009, p. 3). 48 Fried and Polt (2000, p. 27). 49 Krell (1977, p. 166). 50 For a discussion about the poetic solution, see Moore (2012, pp. 479–485). We note, however, that it is not clear that it really succeeds in this aim: it would appear that for one who appreciates the poetry/art, being is still an object of intention, and so an object. 51 Pears and McGuinness (1961, p. xxi). 52 McNeill (1998, p. 85). 53 Fried and Polt (2000, p. 82). 54 Rojcewicz and Vallega-Neu (2012, p. 6). 55 Ibid., p. 58. 56 Ibid., p. 59. 57 Ibid., p. 26.

306  Filippo Casati and Graham Priest 8 Ibid., p. 80. 5 59 Ibid. 60 Ibid. 61 Ibid., pp. 59–60. 62 Gregory and Unna (2009, p. 6). 63 McNeill and Powell (2015, p. 15). 64 Deutsch and Barton (1968, p. 137). By ‘logic’, he of course means Aristotelian logic. 65 Rojcewicz and Vallega-Neu (2012, p. 62). 66 Macquarrie and Robinson (1962, p. 208). 67 Rojcewicz and Vallega-Neu (2012, p. 64). 68 Ibid., p. 29. 69 Ibid., p. 30. 70 Ibid., p. 64. 71 Though, we note, this is partly because interpreters often insist on trying to find interpretations which square with the Principle of Non-Contradiction. Of course this generates a quite needless obscurity. 72 Krell (1977, p. 110). And again: ‘Only because the question ‘What is Metaphysics?’ thinks from the beginning of the climbing above, the transcendence, the Being of being, can it think of the negative of being, of that nothingness which just as originally is identical with Being’. Kluback and Wilde (1959), p. 101. 73 Krell (l977, p. 105). 74 Ibid., p. 106. 75 Ibid., pp. 98 f. 76 Rojcewicz and Vallega-Neu (2012, p. 194 f.). It is also important to acknowledge that some interpreters disagree with the idea that Heidegger characterizes nothingness as the absence of everything (cf. Witherspoon (2002), Severino (1994). However, this is hard to believe because Heidegger himself endorses such a characterization of nothingness in many different essays. For instance, see Mitchell and Raffoul (2012), and Krell (1991). 77 See Chan (1963, pp. 336–337). 78 Lynn (1999, p. 30). 79 Thus, the index of one of his major works, Religion and Nothingness (Van Bragt (1982)), contains 15 references to Dōgen, and 11 to Heidegger. 80 Van Bragt (1982, p. 97). 81 ‘Śūnyata (emptiness)’, is another Mahāyāna term which may be used to refer to ultimate reality. 82 Van Bragt (1982, p. 124). 83 This is an allusion to the Diamond Sutra, which repeats many times statements of the form: x is not x; therefore it is x. 84 Van Bragt (1982, p. 180). 85 Ibid., p. 181.

Works Cited Abe, Masao. 1992. A Study of Dōgen: His Philosophy and Religion, edited by Steven Heine. Albany, NY: SUNY Press. Casati, Filippo. 2016. Being. A Dialetheic Interpretation of the Late Heidegger. PhD Thesis. University of St Andrews. Casati, Filippo and Michael Wheeler. 2016. “The Recent Engagement between Analytic Philosophy and Heideggerian Thought: Metaphysics and Mind.” Philosophy Compass. 9: 486–498. Caputo, John D. 1986. The Mystical Element in Heidegger’s Thought. New York, NY: Fordham University Press.

Heidegger and Dō gen on the Ineffable  307 Chan, W.T. ed. and trans. 1963. A Source Book in Chinese Philosophy. Princeton: Princeton University Press. Cimino, Antonio. 2005. Ontologia, storia, temporalità. Heidegger, Platone e l’essenza della filosofia. Pisa: ETS. Cleary, Timothy. trans. 1986. Sōbōgenzō: Zen Essays by Dōgen. Honolulu, HI: University of Hawaii Press. Conze, Edward. trans. 1973. The Perfection of Wisdom in Eight Thousand Lines and its Verse Summary. Delhi: Sri Satguru Publications. Deguchi, Yasuo, Jay L. Garfield and Graham Priest. 2013. “A Mountain by Any Other Name: a Response to Koji Tanaka.” Philosophy East & West. 63: 335–343. Deutsch, V. and Barton, W.B. trans. 1968. What is a Thing? London: Gateway Editors. Fine, Kit. 1975. “Vagueness, Truth and Logic.” Synthese. 30: 265–300. Fried, Gregory and Richard Polt. trans. 2000. Introduction to Metaphysics. New Haven, CT: Yale University Press. Fritsch, M. and G.A. Gosetti-Ferencei. trans. 2010. The Phenomenology of Religious Life. Bloomington: Indiana University Press. Garfield, Jay L. trans. 1995. The Fundamental Wisdom of the Middle Way. New York, NY: Oxford University Press. ———. 2002. “Sounds of Silence: Ineffability and the Limits of Language in Madhyamaka and Yogācāra.” In Empty Words (pp. 170–186). Oxford: Oxford University Press. Gregory, W.T. and Y. Unna. trans. 2009. Logic as the Question Concerning the Essence of Language. Albany: SUNY Press. Heine, S. trans. 2009. “‘Dōgen’s Shōbōgenzō Fascicles ‘Katto’ and ‘Ōsukasendaba’,” in Buddhist Philosophy: Essential Readings, edited by W. Edelglass and J. Garfield (pp. 149–158). Oxford: Oxford University Press. Hershock, P. 2015. “Chan Buddhism,” in Stanford Encyclopedia of Philosophy, edited by E. Zalta. http://plato.stanford.edu/entries/buddhismchan/ Kassor, Constance. 2013. “Is Gorampa’s ‘Freedom from Conceptual Proliferations’ Dialetheist? A Response to Garfield, Priest, and Tillemans.” Philosophy East and West. 63: 399–410. Kluback, W. and Wilde, T. trans. 1959. The Question of Being. Harrisonberg, VA: Vision. Krell, David Farrell. ed. 1977. Martin Heidegger: Basic Writings. New York, NY: Harper and Row. Krell, David Farrell. ed. 1991. Nietzsche. Volume III: The Will to Power as Knowledge and Volume IV: Nihilism. New York, NY: Harper Collins Publishers. Lilly, R. trans. 1991. The Principle of Reason. Bloomington: Indiana University Press. Lynn, R.J. 1999. The Classic of the Way and Virtue: a new translation of the Tao-te Ching of Laozi as interpreted by Wang Bi. New York, NY: Columbia University Press. Macquarrie, John and Edward Robinson. trans. 1962. Being and Time. Oxford: Basil Blackwell. May, Reinhard. 2005. Heidegger’s Hidden Sources: East-Asian Influences on his Work. London: Routledge. McNeill, Will. ed. 1998. Pathmarks. Cambridge: Cambridge University Press. McNeill, William and Jeffrey Powell. trans. 2015. The History of Beyng. Bloomington: Indiana University Press.

308  Filippo Casati and Graham Priest Miller A.V. trans. 1969. Hegel’s Science of Logic. London: Allen and Unwin. Mitchell, A. and Raffoul, F. trans. 2012. Four Seminars. Bloomington: Indiana University Press. Moore, Adrian W. (2012). The Evolution of Modern Metaphysics. Cambridge: Cambridge University Press. Narbonne, J-M. 2001. Henologie, Ontologie et Ereignis (Plotin—Proclus— Heidegger). Paris: Les Belles Lettres. Pears, D.F. and McGuinness, B.F. trans. 1961. Tractatus Logico-Philosophicus. London: Routledge and Kegan Paul. Priest, Graham. 2002. Beyond the Limits of Thought, 2nd edn. Oxford: Oxford University Press. Rojcewicz, R. trans. 2007. Basic Concepts of Ancient Philosophy. Bloomington: Indiana University Press. Rojcewicz, R. and Schuwer, A. trans. 2003. Plato’s Sophist. Bloomington: Indiana University Press. Rojcewicz, R. and Vallega-Neu, D. trans. 2012. Contributions to Philosophy. Bloomington: Indiana University Press. Severino, Emanuele. 1994. Heidegger e la metafisica. Milano: Adelphi. Siderits, Mark. 2007. Buddhism as Philosophy. Aldershot: Ashgate. Stambaugh, Joan. trans. 1985. Schelling’s Treatise: On Essence Human Freedom. Columbus, OH: Ohio University Press. Tanahashi, Kazuaki. trans. 1985. Moon in a Dewdrop: Writings of Zen Master Dōgen. San Francisco, CA: North Point Press. Thurman, Robert. 2014. The Holy Teachings of Vimalakīrti. University Park, PA: Pennsylvania State University. Tye, Michael. 1994. “Sorites Paradoxes and the Semantics of Vagueness.” Philosophical Perspectives. 8: 189–206. Williams, Paul. 2009. Mahāyāna Buddhism: The Doctrinal Foundations, 2nd edn. London: Routledge. Witherspoon, Edward. 2002. “Logic and the Inexpressible in Frege and Heidegger.” Journal of History of Philosophy. 40: 89–113. Van Bragt, Jan. trans. 1982. Religion and Nothingness. Berkeley: University of California Press

15 The Nietzschean Bodhisattva— Passionately Navigating Indeterminacy George Wrisley

While others have written important works that put Buddhism and Nietzsche in conversation, and others still have promoted a socially conscious Buddhism, I have not found works that recognize the potential of Nietzschean thought to invigorate the Zen Buddhist conception and practice of the Bodhisattva Ideal; the Bodhisattva’s goal being to awaken everyone before themselves.1 As Dale S. Wright notes, Nietzsche has important critical insights concerning the valorizing of a transcendent world and the degradation of this world in the history of Western (religious) values. We can also see Mahayana Buddhism as critical of the renunciation of this world— e.g., through its identification of the “world” of enlightenment2 with that of delusion and its emphasis on compassionate activity in the world. Despite these admirable worldly ideals, however: We see very few images of lives embodying this abstract concern [for others] in practice; few proposals for institutions or sociopolitical orders that really do care for the poor, underprivileged, and those who are suffering. . . . Although Mahayana images of nirvana were crafted to discourage thinking of the ultimate goal as the extinction of finite life, for the most part Mahayana monks continued to practice as though it was.3 Concomitant with this is the ease with which Zen Buddhism can fall into a kind of quietism through the oft seen (partial) description of enlightenment as a state of nonjudgmental acceptance of each moment as it is.4 In response to these sorts of concerns, this chapter seeks to develop the idea of a “Nietzschean Bodhisattva.” Doing so is an endorsement of Wright’s further claim that, “the history of Buddhism is a history of lineages of successive insights and a history of the unfolding of new possibilities for what true excellence in human life might entail.”5 In other words, what enlightenment means is not fixed or static but develops in new ways in response to newly arising conditions. The “nature” of enlightenment is, we should say, itself empty, i.e., impermanent and interdependent with causes and conditions.

310  George Wrisley We will construct our image of the Nietzschean Bodhisattva by looking at Nietzsche’s conception of the higher type of individual in the context of Eihei Dōgen’s Zen and what we can view as its three central indeterminacies— namely, self/other, pain/suffering, and delusion/ enlightenment. It is my contention that a Mahayana Buddhist perspective can fruitfully appropriate aspects of Nietzsche’s conception of the higher type and see the Bodhisattva as a “higher type” passionately engaged in the world in an effort not only to “awaken” all others but to creatively embody “new” values and ways of skillfully alleviating the suffering of others, including the (systemically) oppressed.

The Zen Buddhist Indeterminacies The idea of indeterminacy invites the question of whether it itself is indeterminate. If the base sense of the indeterminate is something like, “not exactly known, established, or defined,”6 do we have equivalent uses of that basic sense in the case of quantum indeterminacy (the inability to determine both the momentum and position of a particle, where if you determine one you rule out determining the other), the example in Wittgenstein’s Philosophical Investigations where the meaning of an ostensive gesture, e.g., pointing, is indeterminate outside a background languagegame, or, to take another example Wittgenstein uses, the Duck–Rabbit illustration of aspect seeing? In this last example, whether it is a duck or a rabbit is indeterminate until it is made determinate by seeing it as one or the other. Regardless of whether these are all the same sort of indeterminacy, they all allow us to see a certain kind of determinacy as the activity of making determinate that which is indeterminate. In quantum mechanics, making one aspect determinate necessitates the other aspect remaining indeterminate. With the ostensive gesture, making one meaning determinate does not make other meanings indeterminate, it simply rules them out. In the case of the Duck–Rabbit illustration, making one aspect determinate does not make the other aspect indeterminate, it simply leaves it there but not “chosen.” Importantly, however, in the Duck–Rabbit illustration, one may see only one aspect, thinking that that is all there is. However, one can “learn” to see both aspects and to shift back and forth between them, and thereby to fruitfully navigate the indeterminacy. In a certain sense, we can see both the concept of the Bodhisattva and the Nietzschean higher type as indeterminate. For, despite all that has been said about the Bodhisattva in Dōgen’s and others’ writings, what the Bodhisattva’s practice comes to needs to be made determinate afresh in new cultural contexts. As Wright stresses: neither Buddhist philosophy nor contemporary standards of thinking would justify Buddhists today continuing to assume as many traditional

The Nietzschean Bodhisattva  311 Buddhists have, that enlightenment is a preexisting human ideal that is fixed and unchanging for all human beings in all times.7 In what follows, I want to use a reading of Nietzsche’s conception of the higher type to make determinate a way of thinking of the Bodhisattva Ideal, in particular, in Dōgen’s Zen. The goal is to invigorate and transcend any quietistic aspects of the Bodhisattva Ideal by way of Nietzsche’s passionate engagement with life; at the same time, this will shed new light on Nietzsche’s project of the revaluation of values, primarily those of suffering and compassion. To be clear, the result is not one I would expect Nietzsche to endorse;8 that is, while I am taking up large “chunks” of Nietzsche’s philosophy, it is not my intent to make my use of them consistent with everything he says, either in any given work or his oeuvre more generally.9 On the other hand, it is my intention to make the position I present here as consistent as possible with Dōgen’s conception of the Bodhisattva Ideal.

Self and Other The indeterminacy of self and other requires entering the mountains and deep waters of the Buddhist concept of no-self. There have been various approaches to understanding and defending the idea of no-self.10 Dōgen does not defend it so much as presuppose it, focusing on expressing it, elaborating on it, and getting the reader/listener to realize it through his writings, Dharma talks, and ultimately the creation of a “new” kind of Zen Buddhism.11 The issue of no-self is difficult because it concerns the self both over time and at any particular time. While we sometimes acknowledge in the West that we are not the same person we were, say, ten years ago, we still habitually conceptualize ourselves as both self-same over time and separate entities/substances. The latter are “self- contained/determined,” i.e., while they may arise from causes/conditions (including other “things), their identity is separate; moreover, they remain numerically identical over time. Because such tendencies of thought are seemingly so ingrained in “human nature” (especially western conceptions of it), Buddhism pushes back rather radically against them. Thus, we find language that denies outright that there are any selves,12 but this denial of self is of a very particular conception of self—again: one that is a self- contained/determined, persisting entity/ substance. One way of approaching the denial of self in the context of Dōgen’s Zen, and “self” can refer to anything’s purported separate and persisting identity, is that something that we ordinarily think of as a self, say, a tree, is nothing of the sort. In regards to persistence, if we pay attention we will see that every aspect of the tree is impermanent. Further, you will see that every aspect of the tree is what it is only because of every-“thing” else. Consider a

312  George Wrisley square, as it is fairly easy to define; we may think of a square as consisting of a set of individually necessary and jointly sufficient conditions. Those conditions constitute any given square. Take any of them away, and you no longer have a square, in concept or reality. Returning to the tree, we might ask what the set of individually necessary and jointly sufficient conditions are that make it a tree.13 Let’s assume a necessary condition is that it have some kind of photosynthesizing structure. The idea in Buddhism is that we err in seeing the tree’s boundary in terms of such necessary conditions. That is, why stop at those particular structures if the criterion for identifying them is that a tree does not exist if they do not exist? For those very structures themselves do not exist unless other “non-tree” things exist. For example, that which we call “a tree” does not exist unless there is also some sort of soil, water, carbon dioxide, sunlight, etc. So, we might say there is no tree unless there are photosynthesizing structures; but if we say that, why not also say there are no photosynthesizing structures unless there is soil, etc.? And, thus, the tree’s existence is coextensive with all of those other “things.”14 To be clear, no-self is the acknowledgement that every-“thing” is a dynamic process of impermanence and interdependence. And the interdependence of all these “things” is not simply synchronic but also diachronic. This tree with its lowlight conditions exists because that tree is next to it and taller: synchronic-interdependence. But this tree at this moment is what it is because of yesterday’s rain. And so the enfolding is both spatial and temporal. The self of my experience is what it is because of what I experience presently, but that present moment experience is also due to past experiences.15 For example, the pathos of today’s flower is due to having seen the same flower at a loved one’s funeral. Each particular “swallows up,” to use Dōgen’s language, everything else, past and present, and thus conditions, i.e., “spits out,” everything going forward. We must be careful here though, while we habitually get trapped into thinking we persist and are contained between our hats and boots,16 we must not get stuck in the other extreme, thinking that we do not exist at all. We can gain some clarity on this issue by briefly examining Dōgen’s firewood analogy. He writes: Firewood becomes ash and does not become firewood again. Yet, do not suppose that the ash is after and the firewood before. Understand the firewood abides in its condition as firewood, which fully includes before and after, while it is independent of before and after. Ash abides in its condition of ash, which fully includes before and after. Just as firewood does not become firewood after it is ash, you do not return to birth after death.17 The not returning to birth after death does not simply refer to what we ordinarily call bodily death at the end of a life, but rather the moment to moment

The Nietzschean Bodhisattva  313 birth and death we undergo. At the moment of firewood we have discreteness— firewood—but that discreteness while independent of before and after simultaneously contains before and after, and all else. Before, as the conditions that gave rise to that moment. After, as what is conditioned (burned) firewood. All else, as the firewood is both diachronically and synchronically conditioned, interdependent with the rest of the world. Hence, Dōgen writes: “there are myriad forms and hundreds of grasses [all things] throughout the entire Earth, and yet each form of grass and each form itself is the entire Earth.”18 And, thus, concerning whether a whole person exists at any given moment, m, we have to use one of Dōgen’s ways of expressing this nonduality of all moments synchronically and diachronically: “although not one, not different; although not different, not the same; although not the same, not many.”19 Thus, each moment is empty of substantial, independent existence, but nothing is lost.20 Each individual moment contains all the rest while being independent of all the rest. We thus find here a unique kind of indeterminacy, as each “thing” at each moment is both discrete and all encompassing. However, again, from the perspective of Zen, we must not fall into the trap of collapsing this indeterminacy in one direction or the other, at least not completely. Thus we have Shohaku Okumura’s explaining that we have to express both “sides” of reality in a single action, and Nishiari Bokusan’s saying, “There is a point in which you jump off both form and emptiness, and do not abide there.”21 Here form means particularity, and emptiness the transitory and interdependent nature of that particularity. The enlightened activity of a Buddha is one of expressing both sides in each particular action one performs; this is jumping off both sides. In practical terms this comes to the wisdom of being able to navigate this paradoxical affirmation and denial of particularity, of self and other. In this way the indeterminacy is taken up and simultaneously transcended.

Pain and Suffering A problem arises if we think of suffering22 and pain as synonymous/coextensive in the context of Buddhism’s claimed ability to give us the path to the cessation of suffering. It is hard to know what it might mean for the human organism not to experience pain. This problem is resolved, however, when we distinguish between suffering and pain, as the Buddha does, e.g., in the Sallatha Sutta.23 Thus, we can say that while pain is inevitable, suffering is optional.24 One suffers the physical pain of, say, a toothache, only when one responds to that pain in a particular way; hence, “When touched with a feeling of pain, the uninstructed run-of-the-mill person sorrows, grieves, & laments, beats his breast, becomes distraught. So he feels two pains, physical & mental.”25 Thus, it is not that an enlightened person does not experience pain; rather a Buddha remains unattached and experiences both themselves and pain as non-substantial, i.e., as continually changing and as what they are only in relation to other things.

314  George Wrisley Those other things are manifold; however, an example is one’s past experiences with similar pains. If one has had a toothache in the past and that toothache lasted five days and kept one from sleeping, then at the onset of a new toothache, one will experience it as suffering if one projects one’s past experiences onto present experience in a way that fails to remain open to the ever-changing and interdependent nature of the world/experience. However, if one lets go of that expectation, if one recognizes that each situation is truly different, then there is simply the pain, not the pain extended forward into time with all its imagined consequences.26 While we might create a variety of taxonomies of pain, one useful one for our purposes is to distinguish between physical and mental/emotional pain. Again, whether any particular pain is suffering depends upon our reaction to it. This is fairly straightforward regarding physical pain but is less straightforward regarding emotional pain. This is because it is fairly complicated to tease out the emotional pain from the emotional suffering, e.g., of losing a loved one. If what makes the difference is a matter of our reaction and that reaction is largely a matter of attachment, then we must get clear about what love is with and without attachment. Much more needs to be said about these complications, but this brief outline of the issue must suffice for here.27 The central point for us is that we can see any particular moment of pain in itself as indeterminate regarding whether it amounts to suffering.

Delusion and Enlightenment “It is axiomatic in Zen Buddhism that delusion and enlightenment constitute a nondual unity. . . .” The difficulty, of course, is how to understand this unity, for as Hee-Jin Kim further notes, “the interface of delusion and enlightenment in their dynamic, nondual unity is extremely complex, elusive, and ambiguous.”28 We have already come across this “nondual unity” in our discussion of the nonduality of self and other. Where Okumura notes that we must express both sides of reality, unity and difference,29 Kim above writes of “two foci.” Whether enlightened or deluded, one navigates the “same” reality: “Delusion and enlightenment differ from one another perspectivally, are never metaphysical opposites. . ., and are both temporal, coextensive, and coeternal as ongoing salvic processes.”30 As Dōgen puts it, “Those who greatly enlighten delusion are Buddhas. Those who are greatly lost in enlightenment are sentient beings.”31 Delusion is where Buddhas “operate,” and that which “is enlightenment,” i.e., reality’s impermanence and interdependence, is what sentient beings are lost in (deluded about). Kim: “enlightenment consists not so much in replacing as in dealing with or ‘negotiating’ delusion.”32 While free of delusion a Buddha nevertheless moves about “in it.” There is a problematic tendency, especially in Zen Buddhism, to overly privilege equality and non-discrimination, as though enlightenment meant

The Nietzschean Bodhisattva  315 escaping delusion/suffering by way of seeing all things as equal and one, free of concepts, conceptions, and distinctions. Contrary to this, the nonduality that is delusion/enlightenment does not fuzz out distinctions and differences, much less the need to differentiate. I follow Kim’s reading33 of this, namely that the point is that we do not realize and actualize emptiness appropriately if we do not make note of differences (one of the two foci), if we do not take the focus of form (delusion) seriously. We cannot, then, properly realize enlightenment without discriminating, without weighing differences, without lingering in delusion while being nevertheless free of delusion. A further aspect of the interpenetration of delusion and enlightenment is that a Buddha enacts enlightenment within delusion, where enacting is meant to highlight that enlightenment is not so much an epistemological state of mind as it is an embodied practice that implicates the mind.34 The deluded sentient being collapses the pain/suffering indeterminacy onto the side of suffering, whereas the enlightened Buddha collapses it onto the side of “mere” pain. Here the enactment is a matter of realized, habituated, and embodied responses to the reality of pain: a Buddha realizes the transitory and interdependent nature of pain, themselves, and the world, and, as we saw earlier, is not “stuck” in either its form or its emptiness; by contrast, a deluded sentient being “lost in enlightenment” cannot help but linger on the “side” of self, delusion, and suffering, experiencing and enacting themselves as discrete, separate selves who are the unique owners of their pain and, therefore, also of suffering. Centrally, in Dōgen’s Zen, enlightenment/Nirvana is not some otherworldly destination. However, neither is it straightforwardly the result of practice. It is not that one has to, say, meditate for years or lifetimes and only then, and maybe not even then, achieve insight into the nature of reality in a way that is supposed to constitute enlightenment.35 Rather, again, enlightenment is something that a Buddha enacts. For Dōgen, this idea is expressed in his talk of the oneness of practice and enlightenment.36 As soon as one sincerely practices, one is already “there,” but we must not understand practice to mean merely to sit in Zazen. One may take the Buddhaform, i.e., sit cross-legged in Zazen and nevertheless not really be practicing because, for example, one does not have the right intention; further, one may practice enlightenment off the cushion, i.e., zazen is not merely something one practices on the cushion. Thus, every activity, whether on the cushion or not is to become zazen.

The Nietzschean Higher Type Nietzsche writes a great deal on the noble, higher types, and their creativity and genius. I want “merely” to focus on what I take to be a few paradigmatic passages so as to lay out in broad strokes key features of his understanding of these terms. Despite his bellicose rhetoric, writers and artists tend to be Nietzsche’s primary examples of higher types—Johann von

316  George Wrisley Goethe and Ludwig van Beethoven being two of his paradigmatic examples. In a lengthy passage from Twilight of the Idols, Nietzsche describes Goethe’s greatness.37 From that passage we can innumerate the following points as descriptive of the higher type. Higher types: [1] take on as much responsibility and as many projects as they can; [2] they create themselves, overcoming/ transcending their past selves; [3] they are tolerant and magnanimous, not because of some pressure from without but because their strength affords them a certain “kindness” and imperturbability; [4] they are joyful in their ability to affirm what is, not condemning suffering or what others may experience as undesirable. While we might be able to extract any number of other characteristics from Nietzsche’s writings, a further central one for our purposes is, [5]: The noble type of man experiences itself as determining values; it does not need approval; it judges, “what is harmful to me is harmful in itself”; it knows itself to be that which first accords honor to things; it is value-creating.38 The projects of the higher type are like those of Nietzsche, Goethe, and Beethoven—they change the world through their creative endeavors, including the “creation” of value. In what follows, I want to try to describe more than defend what I take is a part of Nietzsche’s position regarding value “creation.” Consider his discussion of master vs. slave morality. In considering these differing value systems, Nietzsche evaluates them in relation to their ability to produce humans capable of great cultural achievements. The slave morality for Nietzsche is likely to result in ignoble productions. Thus, it is not a question for Nietzsche as to which system of values corresponds to some kind of Platonic ideal of true virtue; rather, as one might suspect, his (r)evaluation of values depends on his holding something else as of ultimate value, e.g., cultural productions. And, thus, we might see his revaluation and creation of values, not so much as a literal creation of values ex nihilo, whatever that might mean, but rather as a kind of “re-describing” or “reorienting” in regard to what we should value. Nietzsche thinks that, largely due to ascetic religions, we find Western and Eastern civilizations have come to value the wrong things and, importantly, have improperly devalued suffering. Therefore, value creation is a matter of “creating” “new” ways of deeming, and living out, value. The higher individual, then, “creates values” by way of what they deem valuable: “This is important and worthy of my attention and efforts!” Not just anything goes, however. This “deeming” of the higher type must come from strength, not weakness, from an overflowing of their being.39 And Nietzsche makes clear that the higher type is the one who shows true forms of magnanimity and compassion.40 The values created by the higher type are not those of a scoundrel.

The Nietzschean Bodhisattva  317 Consider Goethe; his life is an example of the kind of values he deemed worthy of pursuing and he helped to shape the values of the German people, e.g., through his creative works, the characters he created in his writings, the kinds of lives they lived, what was important to them, etc. Nietzsche presumably sees himself as a similar sort of transformative figure, given his confrontation with suffering and his “creation of values” through his writings. Central to his project is his desire to “revalue” suffering, which he sees as having a long history of being devalued and condemned as the worst thing, that which must be avoided or mitigated (via traditionally conceived das Mitleid/pity/compassion) at all costs and as quickly as possible. Importantly, neither Goethe nor Nietzsche “created” brand new values from scratch, of course. Goethe is responding to, building on, and transforming sentiments and valuations that are already present in his culture, and which have a long history. The same with Nietzsche; moreover, with Nietzsche we see him seemingly harking back to “old” values found in Ancient Greece, for example. However, we might say that even though the values “created” by Goethe and Nietzsche are not wholly new, we can see them as creating values that are new in the sense of the details of their contours and in the ways that they should be lived. For example, what it means to be indomitable in the context of Homer’s Iliad and in the context of Nietzsche’s life are very different in regard to how that way of being is expressed, what it demands, etc.

The Bodhisattva Ideal We might say that the Bodhisattva is defined by the Bodhisattva’s vow. One translation of which is: Beings are numberless; I vow to awaken them. Delusions are inexhaustible; I vow to transform them. Dharma gates are boundless; I vow to comprehend them. The awakened way is incomparable; I vow to embody it.41 As Kim notes, “These vows are recited, reflected upon, and meditated on by monastics day and night, to such an extent that the lives of monastics are, in essence, the embodiment of vows.”42 The Bodhisattva takes this vow so seriously that she delays “final” enlightenment and returns “birth after birth” to help free sentient beings from suffering. There are two basic senses of Bodhisattva in play with Dōgen and Mahayana Buddhism more generally. That is, there is the Bodhisattva as a way of practicing Buddhism, i.e., the pursuit of liberation for all and of the Mahayana ideal; there is also the Bodhisattva as an “object of faith and devotion.”43 In this latter sense, there is a whole “pantheon.”44 In his fascicle “Avalokiteshvara,”45 Dōgen venerates the mythical Bodhisattva of

318  George Wrisley great compassion, Avalokiteshvara. He is said to have a thousand arms and eyes. So many arms (hands) and eyes are representative of Avalokiteshvara’s ability to extend his “infinite compassion” to all beings.46 Given Dōgen’s identification of Avalokiteshvara as the “parent of all buddhas”47 and given that he is the bodhisattva of great compassion, it is not hard to see why Kim would conclude that, “The essence of the bodhisattva ideal [is] great compassion.” Importantly, Kim continues: [The bodhissatva ideal] was [for Dōgen] the reconciliation of the dualistic opposites of self and nonself, sentient and insentient, Buddhas and sentient beings. . . . The identity of “I” and “you” in thusness [emptiness/Buddha-nature], rather than identity in substance, status, or the like, was the fundamental metaphysical and religious ground of great compassion.48 Acts of compassion—which infuse every action of a Bodhisattva who embodies emptiness through the care and attention to everything done, said, and thought—are the expression of the two sides of reality, form and emptiness, in a single, compassionate action. It is vital to note that compassionate action in Buddhism need not be understood “merely” as a matter of teaching the Dharma, i.e., Buddhist teachings/truths. Consider the social movement of engaged Buddhism. A prominent advocate of engaged Buddhism, Alan Senauke, writes: It is hard to define engaged Buddhism. But I think it has to do with a willingness to see how deeply people suffer; to understand how we have fashioned whole systems of suffering out of gender, race, caste, class, ability, and so on; and to know that interdependently and individually we co-create this suffering. Looking around we plainly see a world at war, a planet in peril. Some days, I call this engaged Buddhism; on other days I think it is just plain Buddhism—walking the Bodhisattva path, embracing the suffering of beings by taking responsibility for them.49 On one hand, from the Buddhist perspective, even if there were no systems of oppression revolving around gender, race, etc., people would still suffer profoundly due, we might say, to their ignorance of the Four Noble Truths, etc. Hence, the emphasis in Buddhism on spreading the Dharma. But as Senauke points out, we cannot ignore the various systems of oppression that compound people’s suffering and confound their ability to mitigate it. In this context we should consider Gary Snyder’s warning: Institutional Buddhism has been conspicuously ready to accept or ignore the inequalities and tyrannies of whatever political system it

The Nietzschean Bodhisattva  319 found itself under. This can be death to Buddhism, because it is death to any meaningful function of compassion. Wisdom without compassion feels no pain.50 Thus, it is vital to Zen that the compassionate activity of the Bodhisattva extends beyond teaching those who are under the boot of the oppressor how to avoid experiencing that heavy boot as suffering. We must not become trapped in the idea that the Dharma/enlightenment is enough in regard to alleviating suffering. If, for example, we look at the history of Buddhism, say in Japan, we can see sexist practices flourishing in the midst of purportedly enlightened monks and monasteries. We might put it so: the Dharma needs the wisdom of feminist scholarship/activism and the latter needs the wisdom of the Dharma.

The Nietzschean Bodhisattva Let’s now consider how, in the context of our three indeterminacies, we can view the Bodhisattva as a Nietzschean higher type. To begin, Nietzsche’s higher/lower type dichotomy is transformed in the context of Zen and our three indeterminacies. For Nietzsche, there is a definitive connection between the higher type as a kind of noble master and the lower type as an ignoble slave. This kind of value-hierarchy, as a normative valuation of people’s worth, is anathema to Zen Buddhism. Instead, in the Zen context, “higher” comes to refer to “enlightened”/“skillful in practice” and “lower” to “being deluded”/“non-skillful in practice.” The Bodhisattva enacts enlightenment, operating within delusion. Those who do not enact enlightenment through the way they live their lives, are lost in delusion. It is not that they are less valuable; rather, they are unskillful in their lives, creating suffering for themselves and others, and are thereby “lower” than the Bodhisattva. Let us look at the five points we earlier emphasized regarding Goethe as a Nietzschean higher type and see how they may be transformed and embraced by the (Nietzschean) Bodhisattva. [1] Responsibility and Suffering As Nietzsche emphasizes, “the greatness of man” stands in relation to “how many things one could bear and take upon himself, how far one could extend his responsibility.”51 We see in the Bodhisattva’s vow how much the Bodhisattva takes upon themselves. That is, by vowing to awaken all beings, to “come back” life after life until that infinitely impossible task is completed, the Bodhisattva hoists upon their back infinite responsibility. While there are a variety of “mythic” Bodhisattvas, and while there is the danger in the monastic setting for the Zen Bodhisattva to be relegated to the monastery and not “directly” engaged

320  George Wrisley with the world, the Nietzschean Bodhisattva leaps off the cushion and engages the world in all of its turmoil and messiness. Thus, for the Nietzschean Bodhisattva, the vow is not merely an “abstract ideal” to ground one’s practice, but is an active call to wade into the deep mud of the world.52 In the context of responsibility and suffering, let us consider further the indeterminacy of pain and suffering. In a powerful passage, Nietzsche writes: You want, if possible—and there is no more insane “if possible”—to abolish suffering. And we? It really seems that we would rather have it higher and worse than ever. Well-being as you understand it—that is no goal, that seems to us an end, a state that soon makes man ridiculous and contemptible—that makes his destruction desirable. The discipline of suffering, of great suffering—do you not know that only this discipline has created all enhancements of man so far?53 While it is true that the Bodhisattva seeks to alleviate the suffering of others and thereby their own, this is not the absurd idea that it is possible to abolish, to alleviate all pain, whatever its form. To live is to experience a complex and variegated spectrum of pain. Enlightenment means not the cessation of pain, but the cessation of suffering that arises from how one responds to pain. Nevertheless, it is true that the Bodhisattva seeks to alleviate suffering and minimize pain, and, thus, they are not going all the way with Nietzsche here in regard to wanting pain/suffering “higher and worse than ever.” Nevertheless, the goal of ending pain is itself marked by a) the recognition that pain/suffering can be valuable teachers and the compassion of the Bodhisattva may well include letting others linger in their suffering/pain as long as it is a skillful way to help them further along the path of awakening, a point well in line with Nietzsche here and elsewhere; and b) the recognition that suffering arises from pain when one is attached to removing the pain. Point a) above engages Nietzsche’s mantra that suffering is necessary for human flourishing. Particularly in the context of the pain/suffering indeterminacy, Buddhism can well acknowledge this necessity. But we can also push back a bit against Nietzsche by pointing out that he is perhaps a bit too epistemically cavalier about the ability of people to properly assess the suffering threshold of another, and perhaps even of oneself. That is, while it is certainly true that pain/suffering are vital in many of the ways Nietzsche emphasizes throughout his writings, regarding his reconceptualization of compassion, he does not seem to acknowledge the epistemic difficulties of knowing when and how much pain/suffering is necessary.54 And here we might assert, too, that it is simply not true that without further ado “what doesn’t kill us makes us stronger.”

The Nietzschean Bodhisattva  321 [2] Self-Creation and Self-Overcoming In Dōgen’s Zen, there is no fixed self. What we, in delusion, label a self is an ever-changing nexus of causes and conditions whose fundamental ontological locus is the specious and ephemeral present moment, one which, as we saw, “contains” the past, present, and future. In this context, the Bodhisattva creates “themselves” through the way they engage in the world. As Dōgen makes clear, e.g., in his two fascicles on continuous practice, there is no point at which one is done “making a Buddha.”55 Regardless of one’s “degree of awakening,” there is always further insight, further realization, i.e., new ways to enact enlightenment (centered around compassion) in the ever-changing context of causes and conditions that one encounters. In this way the Bodhisattva both creates themselves and continually overcomes their self, both the individual self that is this body and mind, and the self that is the entire earth. [3] Tolerant and Magnanimous Due to Their Strength Rather close to Nietzsche’s own thinking, we find in Buddhism, and this is certainly applicable to Dōgen’s Zen, the idea that in order to develop patience and tolerance, one needs to have confronted many enemies.56 Patience and tolerance must be “tested” and ultimately habituated, and the only way to do either is to practice them in response to difficult people and situations, which requires magnanimity and strength. The Bodhisattva who is trying to awaken others through the Dharma, and who seeks to work on social injustices, oppression, etc., needs both patience/tolerance and self-discipline. Consider again, the Bodhisattva’s vow to awaken all beings. This is a task that calls for “infinite” patience, tolerance, and strength in a world that resists awakening and which has forms of oppression built into its social, political, and economic structures. The wider the variety of difficult experiences faced, the broader and deeper the Bodhisattva develops skills and abilities to navigate everevolving situations. [4] Joyful in Their Affirming What Is; Not Condemning Suffering as Undesirable For Nietzsche, joyful affirmation is a central trait of the higher type. In his doctrine of the eternal recurrence, Nietzsche challenges higher types to joyfully affirm what is, i.e., every moment of one’s life, including those that are painful, difficult, etc., even if those moments were to be repeated for eternity. For Nietzsche, only the strong, higher, noble types could make such a radical affirmation of their lives. In this context, suffering, too, is revaluated. One should not condemn it or try to ameliorate it as quickly as

322  George Wrisley possible. Rather, suffering’s instrumental and contributory value in regard to being a higher type must be appreciated and affirmed. Similarly, the Bodhisattva, in their acceptance of how things are and in their appreciation of the transitory and interdependent nature of everything, experiences a kind of joy in their simultaneous identification with and letting go of the world they experience. This does not, of course, mean that the Bodhisattva does not experience any of the suffering they confront as painful. When we bring in the indeterminacy of pain/suffering, we can recognize that the Bodhisattva can be moved by the suffering of others, experiencing it as painful, and yet still find joy therein. Recognizing this upshot of the pain/suffering indeterminacy allows us to finesse the above condition on the higher type; namely, the higher type does not condemn suffering as undesirable. The Nietzschean Bodhisattva is certainly moved to try to bring the suffering of the world to an end, but, again, this is not a blanket or flatfooted condemnation of suffering. For Nietzsche, there is not a clear distinction between pain (der Schmerz) and suffering (das Leiden). For the Bodhisattva, however, there is. While they seek to bring suffering to an end, they do not believe that it is possible to bring pain to an end: even if they may try to minimize the pain of others, since others who are not yet awake tend to experience pain as suffering. As noted above, the Bodhisattva recognizes the need for pain, difficulty, hindrances, challenges, in short, enemies, for their practice; hence, they do not shy away from, nor condemn all pain and suffering. [5] Determining/Creating Values Through Their “Deeming” and Way of Living, Their Being Exemplars How are we to understand the Bodhisattva’s creativity? In what way is the Bodhisattva creative? For Nietzsche, the (r)evaluation/creation of values is guided by an appreciation of cultural productions; for the Bodhisattva, the “creation” of values is guided by the notion of skillful action that alleviates suffering. In this context, I want to suggest that there are at least two important senses of the Bodhisattva’s creativity. The first is that the Bodhisattva goes beyond past Bodhisattvas, making the Bodhisattva path their own. This creativity is needed since no two lives are the same in the challenges, obstacles, people, situations, etc., that they engage and navigate. Secondly, the Bodhisattva must be creative in finding ways to alleviate the suffering of others. While not directed at the Bodhisattva, the following points from bell hooks are applicable to the way in which the Bodhisattva must be creative in their engagement with the world beyond “simply” trying to awaken others. Concerning the suffering of the poor, hooks writes, The poor are not fooled when the privileged offer castoffs and worn-out hand-me-downs as a gesture of “generosity” while buying only the new

The Nietzschean Bodhisattva  323 and best for themselves. This form of charity necessarily often backfires. Embedded in such seemingly “innocent” gestures are mechanisms of condescension and shaming that often assault the psyches of the poor. No doubt that is why so many poor people in our culture regard charitable gestures with suspicion. It is always possible to share resources in ways that enhance rather than devalue the humanity of the poor. It is the task of those who hold greater privilege to create practical strategies, some of which become clearer when we allow ourselves to fully empathize, to give as we would want to be given to.57 The contemporary Bodhisattva must navigate a host of difficult issues particularly in a world that is beset with stark hierarchies and dualities involving race, class, gender, ability, sexual orientation, etc. Such complexities call for a sensitive creativity if they are to be successful and not received as insincere or condescending instances of “good will.” In this passage from hooks, we see the four explicit activities of the Bodhisattva.58 That is: kind speech, generosity, beneficial action, and “identity action.” I take it as fairly clear how the first three would need to be manifested in the context of hooks’ passage. The fourth, “identity action,” is not so obvious; it is the translation of the Japanese dōji 同事, which is a standard translation for samānârthatā, the Bodhisattva virtue of “shared concern,” in the sense of “working together” with others.59 The idea is that one identifies with those one helps through shared concern that goes beyond “merely” giving goods. It is the kind of action that comes from working together with others and seeing the other as oneself.60 In the context of Dōgen’s Zen, this is a way of compassionately expressing both sides of reality, i.e., sameness and difference, affirming both separate individual identities and the whole that working together “forms” in the context of emptiness. We see here the reach of the demands on the Bodhisattva. They do not stand aloof, but creatively work together with those they serve toward a common end. And in this context “create,” i.e., help specify/deem what is of value given the end of alleviating suffering. We might even see this as a revaluation of contemporary (Western) values—ones that underwrite practices of inequality, oppression, and suffering. As Nietzsche sought to overcome the values inherent in the history of religion and “create” “new” ones, the Nietzschean Bodhisattva, in the contemporary (Western) world, seeks to overcome the values that underlie modern systems of oppression, e.g., those of capitalism and white supremacy, and thereby “creates” “new” ways of valuing. These new ways are ones that support not only (Buddhist) enlightenment but autonomy, equality, equity, and justice in a world mired in oppression and inequality. While this chapter has been necessarily programmatic and broad brushed in drawing connections between the Nietzschean higher type and the Buddhist Bodhisattva, in closing we might briefly qualify their alignment by noting that Nietzsche views egalitarian goals as privileging the many over the

324  George Wrisley few, the few who, in the world he desires to see, are to be creators of value and culture. However, it is not clear, given the nature of the indeterminacies we have outlined, why a creative genius could not be a Bodhisattva creating culture and value in the context of their vow to awaken and save all sentient beings, as argued here. However, and importantly, the practice of a Nietzschean Bodhisattva could even involve the creation of art, music, etc., all with soteriological ends. Such value creation would not be that of the “high culture” of Western Europe that Nietzsche privileges, but so be it; “high culture” is no goal for us.

Notes 1 Regarding Nietzsche and Buddhism, see Morrison (2002) and Panaioti (2014). Regarding socially conscious Buddhism, see Senauke (2010). A question that is in the background of the thinking in this essay is: in what sense could a Goethe or a Beethoven, paradigms of Nietzschean higher types, have been a skillful Buddhist? 2 While there are a number of issues with the English word, “enlightenment,” (see Wright [2016, pp. 1–2 and 199–202] for a helpful discussion), I use it to refer to the ultimate goal of Buddhism. “Nirvana” is also used for this, but in my mind it conjures up thoughts of some realm that one goes to, whereas “enlightenment” or “awakening” refers more directly to a state of the person or their way of engaging the world, which I find preferable. 3 Wright (2016, p. 204). Emphasis mine. 4 See Kim (2007, p. 35) for more on this issue. 5 Wright (2016, p. 199). 6 Taken from the Apple IOS New Oxford American Dictionary app. 7 Wright (2016, p. 198). 8 While Nietzsche’s views on Buddhism cannot be adequately summarized here (again, see, e.g., Morrison [2002] and Panaioti [2014]), he had a more favorable opinion of Buddhism than he did Christianity. One reason is that he saw Buddhism as having overcome the vengeful temptations of ressentiment (See, e.g., Nietzsche [1992, Ecce Homo, “Why I am so Wise” §6 and The Antichrist §§20– 23]). Nevertheless, Buddhism was still nihilistic for him; together with Christianity, both were “décadence religions” (Antichrist §20). A central problem with Buddhism, for Nietzsche, is that he saw it as disparaging suffering and having at its heart the wrong kind of Mitleid/pity/compassion (see, e.g., Nietzsche [1992, The Genealogy of Morals, Preface, §5]; on Nietzsche and compassion, see more below). Another problem with Buddhism, for Nietzsche, is that he took the Buddhist desire for “Nirvana” to be a desire for a kind of nothingness (see, e.g., Nietzsche [1992, The Genealogy of Morals, Part I, §6]). While this essay is not intended to be comparative, the reader will be able to discern how Nietzsche’s views on Buddhism, particularly in regard to the two main problems he finds in Buddhism (noted above), are problematic. A key problem is that Nietzsche does not clearly distinguish, as Buddhism can be seen to, pain and suffering (more on this below). 9 Something Nietzsche does not always seem interested in doing himself. And, indeed, we can often see Nietzsche as trying on different, often inconsistent perspectives. 10 See, e.g., Duerlinger (2006) and Siderits (2007). 11 See Heine (2006, chpt. 6) for a discussion relevant to Dōgen’s creation of a new form of Zen.

The Nietzschean Bodhisattva  325 12 See, for example, the Heart Sutra. For a discussion of its history and importance, see Tanahashi (2016). 13 I’m not prepared to deny that trees have essences, but I am skeptical. However, for our purposes, I will simply talk in terms of necessary conditions, leaving aside the question of whether we could specify ones that are jointly sufficient. 14 We might worry that this conflates the necessary conditions for the tree to be a tree with the necessary conditions for that which is a tree to exist in the world— the idea of a tree with what is required for that idea to be instantiated in reality, e.g., logical/conceptual necessity with material/physical necessity. This is certainly an important concern. While we can’t begin to address it properly here, we can simply note that the Zen Buddhist ontology that I attribute to Dōgen radically problematizes such a distinction between an idea’s/concept’s essence and what allows that idea/concept to be actualized. An actual tree is what it is, how it is, only because of past and current causes and conditions (including the seed from which it came, etc.). But even in this interdependence, the tree is still a tree—neither completely individual nor completely “washed out” in interdependence, and thus is its indeterminacy. 15 Calling the nature of present experience, say, of a tree, synchronic is perhaps misleading if we bring in considerations of the time it takes for the light of the tree to reach my eyes, etc. However, it seems to me that despite such complications, there is still value in speaking of the synchronic vs. diachronic conditioning of present experience. 16 To borrow an expression from Walt Whitman’s “Song of Myself.” 17 Dōgen (2012, p. 30). 18 Ibid., p. 105. Interpolation mine. 19 Ibid., p. 451. Compare the philosopher who says a person wholly exists at any given moment and the temporal parts theorists who denies this. 20 See Dōgen’s “The Time Being” fascicle, in Dōgen (2012), for his explicit treatment of this. 21 Bokusan (2013, p. 33). 22 We must be careful when we translate the Sanskrit “dukkha” or the Japanese “ku” into “suffering.” All three of these terms have their own nuances; nevertheless, while “dukkha” can mean a variety of things, including a more general dissatisfaction, “suffering” works well-enough and is readily found in translations of Dōgen’s work. 23 This Sutra is a part of the Theravadan cannon, and though I have not found an instance of Dōgen’s referencing it, I take it that it is the most reasonable way for any form of Buddhism to resolve the difficulty of claiming that Buddhism is a path to the cessation of suffering. 24 This is apparently not from a Buddhist Sutra. A google search indicates that its source is anonymous and not attributable to the Buddha. Nevertheless, it is a nice way of making the point. 25 “Sallatha Sutta: The Arrow.” 26 It is important and interesting to note in this context the recent research on pain and pain perception. Contrary to the reasonable seeming idea that the experience of pain is the simple and direct result of nerves responding to stimulation, there is good reason to believe that the experience of pain is a product of nerve stimulation and the way a subject (and we might say their culture/society) contextualizes and assigns meaning to the pain/kind of pain in question. One example of the “plasticity” of pain is called “catastrophizing.” See, the May 2017 issue of Scientific American Mind on pain, which has a helpful discussion of this phenomenon and much more.

326  George Wrisley 27 I am unfortunately unaware of any explicit treatment of this complication regarding the distinction between pain and suffering in the Buddhist context. However, regarding love and attachment, see Newland (2016). 28 Ibid. pp. 1–2. 29 Okumura (2010, p. 18). 30 Kim (2007, p. 4). 31 Ishida (2010, p. 10). 32 Kim (2007, p. 4). 33 For example, Kim (2007, p. 43). Consider, too, Dōgen and Uchiyama (2005, pp. 38 and 46). 34 By “mind” here, I mean more or less the individual mind of a discrete person; however, I do not mean this to contradict or deny the sense of “mind” that is often at play in Dōgen’s Zen, whereby what is translated as “mind” in English can mean a variety of things. See Kim (2004, pp. 116–125) for a helpful discussion of Dōgen’s views on mind. 35 Kim wonderfully takes apart such a view; see chapter 1 of Kim (2007). None of this negates that there is still a difference between sitting with realization and sitting without; but there is not room here to further elaborate on this complicated subject. 36 See Dōgen (2012, pp. 3–22). 37 Nietzsche (1990, “Expeditions of an Untimely Man,” §49). 38 Nietzsche (1992, BGE, §260). 39 See Nietzsche (1974, §370). And in Nietzsche (1992, BGE, §260) we find a line important in the context of this paper: “the noble human being, too, helps the unfortunate, but not, or almost not, from pity (das Mitleid), but prompted more by an urge begotten by excess of power.” It is interesting to consider how we should understand “power” in the context of the Bodhisattva Ideal. 40 As discussed in note 8 above, one of the greatest tensions between Nietzsche’s work and what we are doing here concerns das Mitleid/pity/compassion (das Mitleid for the rest of this endnote). We will not be able to even begin to do justice to this issue; however, I want to at least note one of the more significant issues. In (1992, BGE, §225), Nietzsche makes the claim that in a person there is both creature and creator. Here Nietzsche seems to say that the creature part needs to suffer in order for the creator part to truly be what it is. Ordinary das Mitleid seeks to alleviate the suffering of the creature part. And here Nietzsche distinguishes another kind of das Mitleid; it is a das Mitleid that laments the negative effects of ordinary das Mitleid—how the latter makes not only the individual but also the species smaller and weaker: more mediocre. The higher, creative types need their suffering; das Mitleid directed at them means turning up the resistance, the difficulty of things, the weight and number of responsibilities— turning up the suffering. Again, while this topic is far too large to address here with adequacy, we can say that Nietzsche’s distinction between two kinds of das Mitleid becomes more problematic when we introduce our distinction between pain and suffering. As will be addressed in the body of the paper, das Mitleid of the bodhisattva is much more nuanced than das Mitleid for the creature part that Nietzsche distinguishes from that for the creator part of a human being. See Reginster (2006) for an excellent discussion of Nietzsche’s revaluation of das Mitleid. For a helpful discussion of the differences between pity and compassion in the contexts of both Buddhism and translating from a language that only has a single word for both, see Conway (2001). 41 Tanahashi and Nagase, (2015, p. 30). 42 Kim and Leighton (2004, p. 204). Emphasis mine. 43 Ibid., p. 204. 44 See, for example, Leighton (2012). 45 Avalokiteshvara is also the central speaker in the Heart Sutra.

The Nietzschean Bodhisattva  327 6 Kim and Leighton(2004, p. 207). 4 47 Dōgen (2012, pp. 397–398). 48 Kim and Leighton(2004, p. 208). We will see below in detail what is meant by reconciling these opposites, including what is meant by the reconciliation of the sentient and insentient. 49 Senauke (2010, p. iii). 50 Snyder (1969, p. 90). 51 Nietzsche (1992, BGE, §212). 52 I do not mean to imply that the practice of the monks in monasteries should be viewed simply as quietistic disengagement with the world outside. For example, the monks in their concentrated practice, maintain the dharma as a kind of storehouse, since it is easy for it to become lost or diffuse in the distraction of the world outside. Further, monasteries often engage their local communities. 53 Nietzsche (1992, BGE, §225). 54 An issue that we cannot here take up, but which would be fruitful to address, is whether our distinction between pain/suffering is capable of undercutting Nietzsche’s concern that suffering is necessary for greatness, and since traditional compassion is seen as the attempt to remove suffering, it should also be seen as interfering with greatness. Reginster, e.g., writes, “Nietzsche defines greatness in terms of power, or the overcoming of resistance, so that there cannot be greatness without suffering” (2006, p. 186). Suffering and resistance are here being equated. However, if we can make sense of resistance in terms of pain and not suffering, then Buddhist compassion need not succumb to this Nietzschean worry. 55 See Dōgen (2012, “Continuous Practice” Parts One and Two). 56 See Shantideva (1998, p. 55). 57 Hooks (2000, pp. 47–48), emphasis mine. I am reminded in this context, too, of Benjamin O. Arah’s emphasis of a point made by Dr. King: “[he] argued that his civil disobedience and philosophy of nonviolent protest were creative forms for direct action in the strategic fight against white segregation and racial injustice.” Arah (2014, p. 285), emphasis mine. 58 See Dōgen (2012, “The Bodhisattva’s Four Methods of Guidance”). 59 I owe these points about “identity action” to a very helpful correspondence with Carl Bielefeldt. 60 Writing in this way, I do not mean to imply that the Nietzschean Bodhisattva cannot be a person experiencing oppression firsthand—they certainly may be. Indeed, Arah, again writing on King’s position: “[King] further explained that the one creative thing that an oppressed person can do has to do with how he approaches his state of oppression or condition. . . . ” (emphasis mine). Of the three choices such a person has, “the ‘third way’ or the choice of nonviolent resistance” (2014, p. 285) is the one, as above, identified by King as creative. The point is not necessarily to say that King was a Nietzschean Bodhisattva, though that is an interesting question, but rather to emphasize the ways creativity may manifest and who can be a Nietzschean Bodhisattva.

Works Cited Anālayo, Bhikkhu. 2015. Compassion and Emptiness in Buddhist Meditation. Cambridge: Windhorse Publications. Arah, B. O. 2014. Socrates, Thoreau, Gandhi, and the Philosopher/Activist Dr. King: Politics of Civil Disobedience and the Ethics of Nonviolent Action. In The Liberatory Thought of Martin Luther King Jr.: Critical Essays on the Philosopher King, edited by R. E. Birt (pp. 275–295). Lanham, MD: Lexington Books.

328  George Wrisley Bokusan, Nishiari. 2013. Commentary on Dogen’s Genjo Koan, translated by Sojun Mel Weitsman and Kazuaki Tanahashi. In Dogen’s Genjo Koan: Three Commentaries. Berkeley, CA: Counterpoint Press. Conway, Jeremiah. 2001. “A Buddhist Critique of Nussbaum’s Account of Compassion.” Philosophy in the Contemporary World. 8(1). Dōgen, Eihei and Kosho Uchiyama. 2005. How to Cook Your Life: From the Zen Kitchen to Enlightenment, translated by T. Wright. Boston, MA: Shambhala. Dōgen, Eihei. 2012. Treasury of the True Dharma Eye: Zen Master Dōgen’s Shobo Genzo, translated by Kazuaki Tanahashi. Boston, MA: Shambhala. Duerlinger, James. 2006. Indian Buddhist Theories of Rersons: Vasubandhus “Refutation of the Theory of a Self.” London: Routledge Curzon. Heine, Steven. 2006. Did Dōgen Go to China? What He Wrote and When He Wrote It. Oxford: Oxford University Press. hooks, bell. 2000. Where We Stand: Class Matters. New York, NY: Routledge. Ishida, Hoyu. 2010. “ ‘Genjôkôan’: Some Literary and Interpretative Problems of Its Translation.” Academic Reports of the University Center for Intercultural Education, the University of Shiga Prefecture. Kim, Hee-Jin. 2004. Eihei Dōgen: Mystical Realist. Boston, MA: Wisdom Publications. Kim, Hee-Jin. 2007. Dōgen on Meditation and Thinking: A Reflection on His View of Zen. Albany: SUNY Press. King, Martin Luther. 2006. A Testament of Hope: The Essential Writings and Speeches of Martin Luther King, Jr, edited by James Melvin Washington. New York, NY: HarperOne, an imprint of HarperCollins. Leighton, Taigen Dan. 2012. Faces of Compassion: Classic Bodhisattva Archetypes and Their Modern Expression: An Introduction to Mahayana Buddhism. Somerville, MA: Wisdom. Leiter, Brian. 2002. Nietzsche on Morality. Routledge, London and New York. Morrison, Robert G. 2002. Nietzsche and Buddhism: A Study in Nihilism and Ironic Affinities. Oxford: Oxford University Press. Newland, G. 2016. A Buddhist Grief Observed. Somerville, MA: Wisdom Publications. Nietzsche, Friedrich. 1974. The Gay Science, with a Prelude in Rhymes and an Appendix of Songs, translated by Walter Kaufmann. New York, NY: Vintage Books. ———. 1990. Twilight of the Idols; The Anti-Christ, translated by R. J. Hollingdale, edited by M. Tanner. London: Penguin. ———. 1992. Basic writings of Nietzsche, translated by W. A. Kaufmann. New York, NY: Modern Library. Okumura, Shohaku. 2010. Realizing Genjōkōan: The Key to Dōgens Shōbōgenzō. Boston, MA: Wisdom Publications. Panaioti, Antoine. 2014. Nietzsche and Buddhist Philosophy. Cambridge: Cambridge University Press. Reginster, Bernard. 2006. The Affirmation of Life: Nietzsche on Overcoming Nihilism. Cambridge, MA: Harvard University Press. Sallatha Sutta: The Arrow. 1997, translated by T. Bhikkhu. Online at July 05, 2017 www.accesstoinsight.org/tipitaka/sn/sn36/sn36.006.than.html Scientific American Mind. 2017. “Special Report: Pain: New Ways to Find Relief without Opioids.” 28(3). Senauke, Alan. 2010. The Bodhisattvas Embrace: Dispatches from Engaged Buddhism’s Front Lines. Berkeley, CA: Clear View Press.

The Nietzschean Bodhisattva  329 Shantideva, Acharya. 1998. A Guide to the Bodhisattavas’ Way of Life, translated by S. Batchelor. Dharamsala: Library of Tibetan works and archives. Siderits, Mark. 2007. Buddhism as Philosophy: An Introduction. Aldershot: Ashgate. Snyder, Gary. 1969. Earth House Hold. New York, NY: New Directions. Tanahashi, Kazuaki and Nagase, M. 2015. Zen Chants: Thirty-five Essential Texts with Commentary. Boston, MA: Shambhala. Tanahashi, Kazuaki. 2016. The Heart Sutra: A Comprehensive Guide to the Classic of Mahayana Buddhism. Boulder: Shambhala. Wright, Dale S. 2016. What Is Buddhist Enlightenment? Oxford: Oxford University Press.

16 Body and Intimate Caring in Confucian Ethics Qingjie James Wang

“Body” is often seen as “indeterminate stuff” in contrast with “the determinate mind” that has served as the very ground of Western philosophy since its beginning. For a number of years now, Confucian scholars in China have been calling for us to “see Chinese philosophy as a philosophy of the body” (Zhang 2008:1–11). They want to examine the multiple bodily dimensions of Chinese philosophy in its understanding of love, care, family, gender, and other somatic characters and symbols. This embrace is likely related to the present wave of Chinese and non-Chinese thinkers advocating various visions of a Confucian renaissance. Among the earliest of these includes then-Harvard professor Tu Weiming, who in 1985 pushed for increased scholarly attention to Confucianism’s unique concept of bodily-based knowledge (ti zhi, 體知) (Tu 1987:98–111). Tu put forth then the famous image of the Confucian concept of ren ai, 仁愛, “humane love,” as concentric circles expanding outward from the bodily self. In Taiwan, Yang Rubin devoted an entire book to Confucian bodily values (Yang 1996:1–26), while in Mainland China Li Zehou has also proposed that Confucian thought possesses ontology of emotion. Likewise, Meng Peiyuan has repeatedly described the essence of Confucian philosophy as an emotion-based philosophy founded on an ontology of concordance between the body and heartmind (Li 1985:7–51; Meng 2002:1–23). Almost two decades ago, I also published a paper that emphasized the bodily nature of the Confucian virtue shu. I tried to show that Confucian ren ai is first of all a corporeal and living human love that involves emotion and desire, and that Confucian ethics are the formative-exemplary ethics of virtue based on moral exemplification. Thus understood, ren, this key concept of Confucian morality should be seen as that rooted in the interaction between body and mind as well as interpersonal emotional transmission (Wang 2004:231–254). I would like to continue this line of thought here and discuss my understanding of traditional Chinese views of the body and the Confucian concept of intimateness 親近 (literally, “closeness in personal relations”) that is closely bound to the former, as well as the potential significance of these ideas for the future understanding of Confucian ethics.

Body and Intimate Caring in Confucian Ethics  331

I. The Concept of Body in Ancient China and Its Philosophy The Chinese term for body, shen ti 身體, is made up of two characters: shen and ti.1 The first of these terms, shen, is defined in the Shuo Wen Lexicon as a pictograph portraying the human bodily self2 (Xu 1963:170). A second meaning of shen arose later by the time of the late-Han Shi Ming dictionary, which connected it to a homophonous shen 伸 meaning “to stretch or to extend” (Liu 1993:9). The latter character of “body,” ti, is explained in the Shuo Wen Lexicon as the amalgamation of the twelve body parts flushing out into the human figure. Its meaning comes from the bone system (Xu 1963:86). These bodily parts are divided into the crown, face and cheeks of the head; shoulders, spine and buttocks of the torso; upper arm, forearm and hands of the arms; and upper leg, lower leg and foot of the legs. Clearly, here ti refers to the body in its entirety and its essential, e.g., totality (quan ti 全體), wholeness (zheng ti整體), the essential or the original (ben ti本體), the real (shi ti實體), trunk (qu ti軀體) as well as to its parts, e.g. limbs (zhi ti 肢體), the ends of the body (mo ti末體), etc. The Shi Ming defines the term similarly, telling us that it includes “bones, flesh, hair and blood” as well as the hierarchical body structure (Liu 1993:9). In its later usage and now, however, shen becomes generally used: a) as a noun referring to the bodies of humans or animals; b) for first-person pronounal references in a manner similar to “self” (i.e., “myself” 自身 or “personally” 親身, and also as in “to lead one’s men in a charge through a way of going first” 身先士卒); and c) as a verb meaning to bear something oneself, carry something out, or extend, as seen in the “Miu Cheng” chapter of the Huainanzi, where we find “carrying out the words of one’s own as a gentleman, that is xin (trustworthiness)” 身君子之言, 信也. The usage of ti, on the other hand: a) primarily refers to the body, either in its entirety or its parts; from this, b) is extended to refer to not only living bodies but also the natural forms of all objects; c) in its conceptual sense, ti refers to ontological grounding, e.g., subject and substance; d) in its abstract sense, ti refers to the form, fashion, style, method, and system of things and situations; and e) as a verb, ti refers to a way of cognition through the bodily feeling or to an application from ontological grounding to practical realization. In terms of philosophy, the general understanding of “body” arises from the traditional Chinese cosmological understanding of the world as an organic

332  Qingjie James Wang whole. That is to say, the Chinese idea of the body in its very essence involves traditional Chinese views of how all bodies—in their entirety and parts, internality and externality, selfness and otherness, individuality and communality, core and limbs—relate to one another within natural and social systems. These relationships include the roles, ordering, and interaction of all things. Such a view of the cosmos is the reason why scholars of Chinese thought so often emphasize that traditional ideas of body are deeply connected with other fundamental elements of traditional pre-Qin Chinese society and social life. These elements include the ritual ceremony, the instilment of reverence, the organic functioning of the natural world (as full of flowing life forces), and moral practice (that is, achieving dao de—virtuosity and virtuousness arising from conformity with the natural ways of the universe), etc. (Yang 1996:1–88). According to Roger Ames, the most common conception of body in Western traditions is that of a vessel, but in Chinese tradition the body is seen as characteristically organic. This organic conception of bodies is extended to see all things as organically involved in the larger unity of the cosmos, where they take part in the continuous transformations of yin and yang with a universal permeability and circulation. Human bodies thereby are no more than a doorway allowing us to enter this realm of constant and universal fluctuation that includes transformations between presence and non-presence as well as life and death. We can also understand this as a single, personal perspective within this world from which we enter the process of transformation and consciously experience its spontaneous generation (Ames 1993:157–177). Only within the world where “the source is inexhaustible, then, can the four parts of the ti (body) be deeply grounded. Only when the wellspring never dries, can the nine sensory openings flow.” Our bodies serve thus as the individual vehicles of energy, ritual order and morality. We physically enter existence. We grow, age, and pass away. That is the natural occurrence and development of human life, and within this process our bodies are part of the larger body of the cosmos. We are connected, extend and transcend outwardly. In this way, a type of causal oscillation or interactive transformation forms between the myriad individual entities of the world as well as between these entities and the larger whole they are all part of. These entities contain intrinsic connections that influence and confluence, complement and complete each other, and are thereby generative. It is precisely within this network of interconnected, multifaceted bodily relations that I can develop ren, “humanity,” fully realizing my potential as a human being.

II. The Four Central Concepts of the Confucian Philosophy of “Bodies” We should point out that these conceptions of bodily self and organismic cosmology are based on an understanding of concordance between people and

Body and Intimate Caring in Confucian Ethics  333 the cosmos and that they play the foundational role in traditional Chinese thought, especially in Confucian understanding of morality. We find these concepts expressed in the theoretical levels of all major thinkers, e.g., Confucius’s ideas of humane love and benevolent government based on familyrooted intimateness and respect; Zi Si 子思 and Mencius’s teachings on good nature of human disposition; Zhang Zai’s 張載 emphasis on seeing all people and all things as organically related through their shared family and therefore world and cosmos; and Wang Yangming’s 王陽明 advocacy of an engaged, active moral conscious as the highest level of existence. Therefore, when we talk today of a “Confucian renaissance” and revival of the root of Confucianism, this traditional concept of body is undoubtedly one of the elements of Confucian thought for which contemporary re-investigation and re-analysis will prove most fruitful. So then, how exactly can we return to the root of the traditional Chinese concept of body and revive them? In my view, we might as well start it by examining the several philosophical concepts or categories implicit in this idea. As we have shown above, the most important concept in the traditional Chinese view of the body is that of shen. I think that this concept has two principal philosophical implications. The first is related to the idea of the bodily self as myself (自身) and own-self (親身), and the other to a form of self-returning (反身) that involves “returning to myself.” In this way, we can see shen as the origin, a starting point as well as an end for all of Confucian philosophy of self and humanity. The Analects tell us, “Filial piety and respect for elder brothers are the root of humanity,” and the Xiaojing 孝經 says, “The hair and skin of the body are the beginning of filial piety.” Accordingly, filial piety and brotherly respect are founded on connections that are based in the bodily self. Shen 身here clearly refers to one’s own natural body. As a natural body, it is also the vibrant, living self that connects us by blood, flesh, and emotions to our families, by its humanity to our communities, and by its existence to the natural world and cosmos. It is precisely for this reason that Confucians laid so much emphasis on the body. Without this natural base and its connections, as well as the natural emotions that are related to them, the entire Confucian project would find itself hollow or meaningless. In addition to emphasizing the body’s centrality to these familial, communal, and cosmological connections, Confucians also see the body as essential in self-reflection (省身) and self-cultivation (修身)3 (Zhang 2013:336–344). When telling us that “being virtuous (为仁) starts from ourselves,” Confucius is referring to just such bodily-aware self-reflective cultivation. His disciple Zengshen曾参, moreover, emphasizes the importance of regularly reflecting on one’s body-self in moral cultivation (Analects 1:4). Mencius is even more explicit, telling us, “The myriad things of the world are all contained within me. Returning to my body-self (shen) makes me sincere, and there is no greater joy than this” (Mencius 7A4). All of these present

334  Qingjie James Wang reflection on and returning to one’s shen as a process for attaining moral virtue and humanity. This practice starts with returning to the root of our hearts and natures, which are also the base of our behavior. All of this is inextricably related to our bodies being part of the larger body of the cosmos. Only through using reflection on our shen to return to our “root” can we become fully realized, i.e., become humane people—the Confucian junzi, or gentleman. The second main concept in the Confucian understanding of body is ti. As mentioned above, ti can refer to both individual and whole bodies, and it is similar to the concept of shen just described. Thus we see a primary meaning of ti as referring to a body in its wholeness, substance and essence. Other uses of this term, however, also include its verbal meaning of “to experience” or “to be involved in,” which was expanded from ti’s more original meanings of the essential substance (本體), the flesh chunk (軀體), and the organic whole(整體). This second meaning of ti therefore focuses on doing things through extending one’s body-self, as seen in the phrase shen ti li xing 身體力行, which might be idiomatically rendered “practice what one preaches.” In this maxim, shen refers to one’s body-self and ti acts as a verb connoting “to experience” and “to expand.” The latter two characters tell us to “do earnestly,” giving us more literally “experience through one’s body-self trying hard to carry something essential out.” Through this aspect of ti, Confucianism developed ethical theories replete with methodology, consciousness, and praxis that see the natural body as the root or ground of both action and intimate relations with others. Tu Weiming’s concept of bodily knowledge or ti zhi 體知, I think, is basically developed on this use of ti. A number of terms in modern Chinese for different forms of experiencing and cognizing the world around us include the character tiyan 體驗 (to experience); tixian 體現(to manifest); tiren 體認(to realize); tihui 體 會 (understanding through personal experiences); ticha 體察 (to observe), as do a number of words for various versions of consideration of others, e.g., titie 體貼(to considerate); tiliang 體諒(understanding of and tolerance for others); tixu 體恤 (to sympathize), etc. As we know, this kind of bodily knowledge as it was developed in the Neo-Confucianism of the Song and Ming dynasties was understood as a “knowledge of virtue” (德性之知) based on the so-called “unity of knowing and doing ”(知行合一). It differentiated from the “knowledge by empirical cognition” (見聞之知), which was viewed as much more external, instrumental, and trivial. In a contemporary setting, such bodily knowledge sits in contrast to conceptions of scientific knowledge in Western theories of knowledge, which is based on sensory experience, logistic reasoning, and deduction. The third concept included in the Confucian understanding of body is qi 親(intimateness). According to Confucian understanding, this intimateness involves aspects of physical nearness connected to the human body. The Guang Ya dictionary 廣雅 defines it as “nearness”(近). One may point

Body and Intimate Caring in Confucian Ethics  335 out that this idea of “nearness” refers to a relationship in space and time. To this astute observation I would emphasize in response that Confucian concepts of space and time differed from the elaborately abstract theories of modern scientific definitions. Confucian space and time were rather firmly embedded in the dynamic relationships and connected to the physical body. The central Confucian ethical concept of ren ai (humane love) is founded on intimateness in human relations. This human intimateness is based on relational ties (especially of family) that are in turn inextricable from our physicality (that is, “flesh and blood” relations). This is why Mencius tells us that ren is based on not only nearness but also closeness in family relations. This idea of intimateness, which exists within the Confucian concept of body, then, not only constitutes the source of ren ai but also is the basis and impetus for extending ren ai outward from the self. It is intimateness that creates the degrees of closeness in relationships among people that call for and structure the gradations of humane love. The fourth concept contained in the Confucian idea of body is that of bodily-based “gradated area” such as “intimate relations” (伦) and “regions”(域) that extend accordingly outward to others. As previously discussed, the term body primarily refers to the physical human (or animal) form. However, in Confucian understanding it is a living entity that is generated and develops within and through interconnectedness with its surrounding world, its engaged community, and especially its intimate family. The gradated nature of these relationships creates various “bodily regions” that extend through bodily relationships. Large and small, these regions extend and shrink through our meaningful connections, associations, and entanglements with each other. The extension of these bodily gradated regions takes place through the permeability and interconnection of things. This relationship between humans and the cosmos, in turn, relies on Confucian conceptions of the shared existential status of all worldly entities and their cooperation as a single macrocosmic body. The classical Chinese vision of the cosmos includes expansive space (yu宇) and the passage of time (zhou宙),4 which together make up the universe (yu zhou 宇宙). Although they do not themselves exist in a strictly defined physical Newtonian space, bodily regions, therefore, nevertheless involve the extension of ourselves into space and time through mutually formative interaction with our communities and environments. As parts of the macrocosmic body, these bodily regions can be on the one hand extended infinitely. This gives Confucian ren ai its capacity to spread to larger communities of all live beings and even to the whole universe. At the same time, due to their grounding in each of the physical bodies, the gradated regions cannot simply launch un-reined through linear time or physically extended space. Grounded in our individual body and personhoods, they are delimited by our physicality and mortality. That is to say, their extension outward remains tied to the limits of our body and life, and is therefore constrained and finite. The physical borders between persons,

336  Qingjie James Wang constraints of our bodies, and chronological succession of generations inherently limit and define each of us. These regions, therefore, can stretch infinitely outward, but only from and through our actual bodily existences and the recognition of the otherness of other entities, bodies, and persons that our bodies inherently produce. All of this also informs and reinforces Confucianism’s very pragmatic approach to life and moral practices, which involve a gradation of love based on closeness and intimateness of relations and regions.

III. The Confucian Idea of Intimate Caring and Its Relation to Universal Love Based on the traditional Chinese conception of body, we can connect the above four basic concepts into a cohesive view of the Confucian teaching of “intimateness” or intimate loving and caring. Clearly, the core of this intimateness lies in the interactive caring and loving, including different types of convergence, influence and communication among individual entities or ti 體.5 These relationships extend outward from and manifest through the body through gradated relations. The most predominant moral values of Confucian ren ai emerge herein. For example, the central virtues of shu and zhong as a Confucian version of the Golden Rule tell us how to involve “extending our loving and caring to others” by making oneself to others an intimate (近) exemplar rather than a universal standard. The Great Learning also teaches us to see self-cultivation as the “root” of virtuousness and the foundation for successful management of family, state and the world. Such real-world, living depictions of “intimateness” are implicit in Confucian teachings of “extending oneself to others” (tui ji ji ren推己及人) and “comparison between one’s heart and other hearts” (jiang xin bi xin 將 心比心), which have major currency in everyday Chinese life. These concepts of “comparing” and “extending” cannot truly occur without bodily intimateness and its interconnected caring and loving. Therefore, they may not be seen as mono-directional coercive “extensions” driven principally by rational knowledge or determinate calculation, but rather as processes of bodily interactive and indeterminate influence, confluence, and extension taking place between mutually convergent heart-minds. In the same way, Confucians speak of exemplification through oneself in moral practice, and not of the issuance of universal laws and rules via divine Revelation or logistic deduction. Such exemplification works through personal immediacy— relational closeness—which formatively influences others by moving, inspiring and influencing their heart-minds. This impels people to see the world from others’ perspectives and understand the world through indeterminate inference, gradually forming exemplifying models that lead to the cultivation of people’s character and the goal of achieving fully prosperous lives. From ancient times through to the present, however, Confucianism’s teaching of carrying out ren ai through intimate bodily relations which are

Body and Intimate Caring in Confucian Ethics  337 by its nature indeterminate has proven great fodder for debate. More than two millennia ago, Mozi’s advocacy of jian ai (love without gradation兼 愛) faced off directly against Confucians’ advocacy of qin ai (graded love 親愛). Today, an increasing number of people criticize the mainstream philosophy of mind’s belittling of and even hostility toward the body, and they have recognized, in Nietzsche’s words, the “enormous mistake” (Nietzsche 1968:131) of this method. Confucianism’s embrace of the role of the indeterminate body has, thus, subsequently attracted increasing levels of attention. It is important, however, to recognize that Confucianism’s “language of the body” is, at its essence, not a modern innovation. People have spoken of, thought about, and understood the body this way since ancient times. This remains true despite the variation in reception and interpretation of these concepts throughout various periods in history, which followed the oscillation of social trends and fluctuations of Confucianism’s social and political status. For example, the understanding of Confucianism’s core values among the Chinese scholars baptized in the May Fourth and New Culture movements do not fundamentally differ from that of contemporary scholars. Moreover, in the depth and breadth of this understanding, I’m afraid today’s scholars may fall far short of our predecessors. However, critical evaluation of these core concepts contrasts sharply between these two generations. My question is: why is there such a great difference between contemporary and past judgmental positions on these values? Moreover, we may recognize that a “language of indeterminate body” isn’t entirely absent in the West, but rather distorted and marginalized by the “language of the determinate mind” that has dominated Western philosophical and cultural discourse since Descartes. However, we should be careful not to be swept up in trends or chase novelty. That is, we shouldn’t allow the popularity of postmodern views and subsequent reevaluation of the potential value of Confucian views of the body to lead us to ignore or overlook the original aspects of the traditional Confucian understanding of body and self— especially certain fundamental shortcomings of some of these aspects. All contemporary interpretations of the traditional Confucian view of the body must respond rationally to historical criticism within the new framework of our contemporary situation. Otherwise, such new interpretations may prove more of a step backward than forward. In contemporary Chinese history, one of the most influential and important criticisms of Confucian bodily-based intimateness comes from a well-known Chinese sociologist Fei Xiaotong 費孝通. In his famous work From the Soil (《鄉土中國》), Fei takes a sociological perspective in comparing the fundamental difference in social structure between traditional Chinese rural society and Western civil society. According to Fei, the basic structure of Western civil society can be described as top-down, a “public arrangement” (團體 格局) beginning in the public sphere and moving toward the private sphere. Traditional Chinese society, on the other hand, was essentially bottom-up,

338  Qingjie James Wang extending from the personal sphere to public life in a “graded arrangement” (差序格局). The Western “public arrangement” relied on a transcendent and universal conception of the divine, while the Chinese “graded arrangement” based itself on the Confucian principles and practices of intimateness and extending love. Fei tells us that the core of Confucianism’s “extending oneself to others”—which arises through personal humane love—centers on one’s self (自我) and one’s own body (自身). Extending outwardly from individual personal experience, this “extension” casts its ripple within the community in which one socially engages. As this ripple spreads further, it becomes weaker and weaker, and finds itself unable to extend beyond the social circles of Chinese society’s traditional “graded arrangement.” Therefore, Fei concludes that traditional Chinese ethics proved unable to find a “universal moral overview. All standards of value were unable to survive outside of the graded system of human relations” Fei 2006:23–25). Fei thus criticizes traditional Confucian morality as bound to domestic social relations and therefore incapable of universalization and calculative determination—that is, “unable to extend beyond the social order.”6 Fei Xiaotong’s criticism of the intimate nature of Confucianism’s humane love is representative of thought influenced by the mainstream of sociology in the early of 20th century. However, I would like to present two points in response to this criticism against the traditional Confucian view of bodily indeterminacy. First, the basic argument this criticism adopts does not fundamentally differ from that of Mozi who advocates for “ungraded love.” Fei could be only seen as adding some elements of contemporary Protestant ethics to this position.7 In my view, this might be the most questionable aspect of his argument. As we know, the idea of universal love in Protestant ethics relies intrinsically on its presupposition of a divine, transcendent being, that is, God, and Mozi’s conception of universal love is also closely related to belief in an extra-worldly God, heavenly will and spirits. Confucianism, on the other hand, focuses rather exclusively on worldly human affairs. In the Analects, Confucius puts forth an attitude of “respecting spirits but keeping them at a distance” that is integral to the consciousness and context within which Confucian “intimateness” and “indeterminacy” works. How would then a non-divine-based or non-religious social structure and life be possible in China? As we have seen, instead of adopting transcendent grounds, Confucianism is interested in how secular love and care is possible among people in a society without reliance on spiritual, extra-worldly, and absolutely determinate powers. We should point out also that this secular cultural and intellectual context persists continuously into China in the present day. In other words, the last two millennia of ideological and social development in China have seen the Confucian idea of “intimateness” absorbed into mainstream Chinese thought. At the very least, this absorption demonstrates the strong influence of pragmatic and historical “reason” embodied in the Confucian concept of indeterminate “intimateness.”

Body and Intimate Caring in Confucian Ethics  339 Second, Fei’s criticism is to a certain extent built on a common and historically established misunderstanding of “intimateness” in Confucian ren ai. This misunderstanding overlooks the essence of the Confucian view of the body and the bodily self. It, therefore, fails to properly grasp the nature of Confucian ethics. As described above, the core of Fei’s criticism claims that the “intimate” nature of Confucian humane love blocks it from “extending beyond the social order.” First of all, Fei supposes the Confucian idea of “extending love” cannot be equated with the modern natural and social sciences idea of extending (i.e., universalizing) a principle. The latter involves logical inference, deduction, and conclusions, and implies very strong aspects of linear and temporal prediction and determination. On the other hand, Confucianism speaks of remaining firmly within one’s situational context and comparing one’s heart with other hearts—transmission and resonance between organismic bodies. This is essentially what Mencius referred to as “being good at extending” (shan tui 善推). Love or “intimateness” for Confucians, then, is closer to a nonlinear, indeterminately diffuse, and bodily-based affection of the heart-mind. This is a visceral caring and understanding. Therefore, at its core, extension of love and care may not require the universalizing “extension” as in the natural sciences, but rather should be seen as a type of resonance and harmony that occurs through interaction, mutual influence, reciprocal guidance, correspondence, and inter-penetrability. In this way, it works though a sort of resonant ripple effect. Rather than being a wave deliberately generated to drive a specific secular purpose, love and care in this conception become the reverberations of a lingering, circuitous orchestral harmony, in which correspondence continuously increases the strength of resonance. This implicates the very foundation of our understanding of Confucian ethics, which is formative rather than normative and operates through exemplification of virtues rather than following categorical rules. While in this Confucian system logical inference or determinate deduction doesn’t work to spread love and care universally, being-morally-moved (道德感 動) or moral affection does. Therefore, if we view Confucian ethics properly as formative and exemplifying, how far it “extends” and whether it can “extend beyond the social order” may no longer be particularly problematic.8 Therefore, Confucian “intimateness” primarily promotes the cultivation of moral virtue of each person. However, as Confucianism developed, it took on more and more public and political aspects of life. If we then attempt to fully politicize and philosophize this idea, decontextualizing and reapplying it within the framework of modern public and political life, it necessarily causes problems of pan-moralism. By pan-­ moralism I mean a certain exaggeration of the moral claim that mistakenly extends the primacy of a particular conception of morality, along with its motivation and justification, into social and political areas. These areas ought to be, rather, governed by particular and standard rules deriving from the given legal system.

340  Qingjie James Wang Therefore, the “graded” nature of intimacy reveals the very nature of Confucian ethics. Here the living, organic, contextual and indeterminate Confucian body and bodily self means that the gradation it produces does not involve absolute and unchanging determination and delimitation, but rather exists within “regions” that are tethered to the body but capable of expanding infinitely outward to care and to love—albeit an ordered, graduated love—all things.9

Notes 1 Among the earliest records of the two characters shen and ti being used together is the line from the Xiao Jing 孝經, “The hair and skin of our bodies are given to us by our parents. The unwillingness to destroy or harm ours bodies is therefore the beginning of filialness” 身體髮膚, 受之父母,不敢毀傷,孝之始也. 2 Additionally, Li Xiaoding, in Collected Explication of Oracle Script defines shen as “Coming from the image of a human with a protruding abdomen, portraying the shape of the body of a human, serving as the an early ideographic character for body. This is what Xu Shen meant in defining shen as “portraying the bodily self of a human.” 從人而隆其腹, 象人有身之形,當是身之象形初字。許君謂‘象 人之身’, 其說是也.” 3 The Confucian concept of“xingshen 省身”form an interesting and highly telling contrast with the concept of “self-reflection” in contemporary Western philosophy.. 4 As in the classic idea found in Shizi 尸子that “the space of the four directions and above and below is called yu; the passing of the past and coming of the present is zhou” 四方上下曰宇,往古來今曰宙. See Zhu Hailei (Zhu 2006:105). 5 That is, between the many ti’s of part (zhi ti肢體) and whole (zheng ti整體), internal (nei內) and external (wai外), self (ji己) and other (yi ti異體), and individual (ge ti個體) and community (qun ti群體). 6 In the original Chinese, tui bu chu xu 推不出序. Zhao Tingyang has also supported this criticism of core Confucian values. According to Zhao, “Family ethics cannot extend to become social ethics, people whose love is based on family relations (qin ai) cannot love those they are not related to. This is the critical problem of Confucianism.” See Zhao Tingyang (Zhao 2009:129–147). 7 As Fei Xiaotong points out, one may find a transcendent Spirit in the concept behind the Western “public arrangement.” Because of this, the West also has “two derivative ideas: that each individual person is equal before God, and that God is just to each person. . . . The individual and personal intimate connections between parents and children are denied here. . . . Each ‘child,’ in addition to his or her own personal father, necessarily has a more important shared ‘heavenly father,’ which creates the public.” In contrast, “graded arrangements” create societies of “networks of innumerable personal relationships” in which “the scope of society extends outward from the self, and the process of extension can take a great diversity of paths, the most fundamental of which is that of family: parents and children, brothers and sisters, the corresponding moral factors of which are filialness and respect for one’s siblings. . . . Another path for this extension is friendship, to which loyalty and trustworthiness correspond.” Fei thereby concludes, “Thus, traditional morality doesn’t seek an universal ethical concept, and all values find themselves unable to survive independent of the graded arrangement of human relations.” See Fei Xiaotong (Fei 2006:23–35). 8 On this point, Shi Yuankang has put forth the insight that modern people believe “moral problems are limited to issues generated by conflicts of benefit

Body and Intimate Caring in Confucian Ethics  341 and harm. . . . The ethical thought of Confucianism is in this way not at all modern. . . . The ethical thought of Confucianism fundamentally sees morality as an activity of the cultivation of character.” See Shi Yuankang (Shi 1998:116–117). 9 An earlier draft of this chapter was written in Chinese. I would like to thank Bobby Carleo who worked once as my research assistant for his help in preparing the English draft of the chapter.

References Ames, Roger. 1993. “The Meaning of Body in Classical Chinese Philosophy,” in Self as Body in Asian Theory and Practice, edited by Thomas Kasulis, Roger Ames, and Wimal Dissanayake. Albany, NY: CUNY Press. Fei, Xiaotong. 2006. From the Soil: The Foundation of Chinese Society, edited by Liu Haoxing. Shanghai: Shanghai Renmin Press. Li, Zehou. 1985. On History of Ancient Chinese Thoughts. Beijing: Renmin Press. Liu, Xi. 1993. Shi Ming釋名, Bi Yuan 畢沅annotated edition, Beijing: International Culture Press. Meng, Peiyuan. 2002. Emotion and Reason. Beijing: Press of Chinese Social Sciences. Nietzsche, Friedrich. 1968. The Will to Power, translated by Walter Kaufmann. New York: Random House. Shi, Yuankang. 1998. “Two Concepts of Morality.” In Paradigm Shift: From Chinese Culture to Modernity? Beijing: Sanlian Publisher. Tu, Weiming. 1987. “On Confucian Bodily-Based Knowledge and Its Implications of Knowledge of Virtue,” in Proceedings of Conference on Confucian Ethics, edited by Liu Shuxian. Singapore: Institute of East Asian Philosophies. Wang, Qingjie. 2004. “The Golden Rule and Confucian Ethics,” in Heidegger and A Hermeneutical Interpretation of Confucianism and Daoism. Beijing: Renmin University Press. Xu, Shen. 1963. Shuo Wen Lexicon. Beijing: Zhonghua Publisher. Yang, Rur-Bin. 1996. The Confucian Concept of the Body. Taipei: Institute of Chinese Literature and Philosophy, Academia Sinica. Zhang, Shuguang. 2013. “The Philosophy of Body: Self-Reflection, Transcendence and Intimateness,” in Body, Gender, Family and Symbols, edited by Zhang Zailin. Xian: Xian Jiaotong University Press. Zhang, Zailin. 2008. Ancient Chinese Philosophy as a Philosophy of the Body. Beijing: Chinese Social Sciences Press. Zhao,Tingyang. 2009. A Study of Bad World—Political Philosophy as the First Philosophy. Beijing: Renmin University Press. Zhu, Hailei. 2006. Annotation of Shizi. Shanghai: Shanghai Guji Press.

17 Indeterminacy in Chinese Thought Spontaneity and the Dao Robert Cummings Neville

Introduction My purpose in this chapter is to introduce the broad strokes of how traditional Chinese philosophy has handled issues of indeterminacy. This is not original research into Chinese texts, but rather is an exhibition of the strategies for handling what many forms of Western philosophy have identified as indeterminacy. Furthermore, I shall not disguise some of my own philosophy, which has learned from both Chinese and Western sources, to interpret indeterminacy. Two broad problematics of indeterminacy obtain in both China and the West. The first is what I call cosmological because it has to do with indeterminacy in how the processes of nature, institutions, and personal life play out. In Plato’s Republic and many other dialogues, the characters debated how to balance or harmonize the many natural, social, and personal factors in play, supposing that it was not absolutely determinate what was going to happen in advance of developing good judgment and choice. Aristotle discussed the indeterminacy of the future when a group of admirals debated whether to go to battle on the morrow (in “On Interpretation,” Chapter 9). Set against this cosmology of partial indeterminacy was the principle articulated by Aristotle and winning favor through much Western (and Muslim) medieval philosophy that all the actuality in an effect must be contained in its antecedent causes with nothing new: ex nihilo, nihil fit. Although many interpretations of this arose, the principle basically said that the present, where choices might be made, and the future are determined by the past; no indeterminacy. With the rise of early modern science, Aristotle’s complicating formal, final, and material causes were eliminated from scientific consideration, and efficient causes alone were allowed to explain, resulting in rigid “determinism.” Kant went so far as to say, in the Critique of Pure Reason, especially the “Second Analogy,” that determinism must be presupposed by science so as to distinguish between the objective order and the subjective order of representations. This scientific determinism was qualified in the 19th century with the rise of statistical causal explanations: a population (say, of gas molecules) is determined even though its members (the

Indeterminacy in Chinese Thought  343 individual molecules) are not. I shall trace down the Chinese alternatives to this Western story of indeterminacy. The other problematic of indeterminacy I call ontological because it has to do with how a determinate, or at least partly determinate and partly indeterminate, cosmos arises from a simple, undifferentiated, indeterminate ground. Plato said in The Republic that the Form of the Good gives rise to or creates everything knowable in any sense and all kinds of knowing in any determinate sense. But he did not elaborate that point. In the Parmenides he showed just how complicated this notion is when run through the changes of the problem of the one and the many.1 Aristotle said little or nothing about this. But Plotinus in the Enneads said that the One or Being is beyond all determination and overflows to give rise to it. The One in itself is indeterminate and it gives rise to the Dyad, and hence to all else, which is where determinateness of any sort lies. Versions of this “infinite fullness of being contracting itself to intelligibility or determinateness” developed in pagan, Christian, Jewish, and Muslim Neo-Platonic philosophy. Though influenced by Neo-Platonism, Thomas Aquinas went on a slightly different tack to modify Aristotle’s idea of “being in act,” as in finite actualities and actualizations, into the idea of God as the Pure Act of To Be, infinite and simple. Aquinas’s God creates finite things by putting limitations on the infinite (i.e., indeterminate) Act of To Be. However, those things are separate in the created world but are indistinguishable within the divine essence, which is Pure Act with no potentialities or delimitations. The Neo-Platonic and Thomistic conceptions construe God as an infinite fullness that is somehow contracted or delimited in the creation of the determinate world, moving from ontological indetermination to the ontological being of cosmological determinateness. These positions called themselves creation ex nihilo, meaning creation from no determinate thing. In contrast, Descartes denied the fullness of being interpretation of the indeterminate and said that all creation of determinate things, including intelligibility, is positive creation of something from absolutely no thing, not even fullness. Finite things, for Descartes, are wholly positive realities that can be known (with the proper method) by the natural light of reason. The positive realities are entirely novel, not made from antecedent actuality, infinite or finite. Descartes’s general position was followed by Paul Tillich (and me), among others.2 For all these ontological positions and their variants, any determinacy is grounded in, created by, or caused by the indeterminate. Another way of saying this is that the complex arises from and is to be explained by the simple, somehow or another. These positions of creation ex nihilo stand in contrast with positions that say the most fundamental source of finite things is itself determinate and rational, such as Aristotle’s Thought Thinking Itself, Leibniz’s God finding the best harmony of compossibilities in order to have sufficient reason for creating, and Hartshorne’s God of necessary metaphysical first principles.3 I shall discuss some Chinese alternatives to these strategies for explaining how the determinate arises from the indeterminate.

344  Robert Cummings Neville

Cosmological Indeterminacy The aboriginal Chinese sensibility about the cosmos is that it consists of changes.4 To find anything like the notion of things as substances is difficult if not impossible in Chinese philosophy, literature, or art. The cosmos is composed of a universal stuff, qi (pronounced “chi”) that has different scales of changes, from the rough changes of mountains that seem very solid and extremely slow to change, through the swifter and more complicated changes of biological nature to the extremely subtle changes of highly complex and interactive changes of human thinking, even the speediest or most tranquil thinking of the sages. Mind–body dualism is almost impossible to imagine in the Chinese philosophical sensibility. The closest is a problematic of inner personal development versus outer action in the world, but even here the dominant positions have stressed the continuity between thought and action.5 The fundamental units of changes in qi are yin and yang.6 Yang is a movement of extension or expansion. When it reaches its limits or end of its resources, it stops and begins to contract. Yin is a movement of contraction, consolidation, returning to a matrix of resources. No such thing as a static yin-yang state of affairs exists, in Chinese thinking; rather only there are only increasing and decreasing yang and increasing and decreasing yin. The concepts of yin and yang have great metaphoric reach. Originally, yin meant the north side of the mountain, which in China’s northern hemisphere location was associated with winter, cold water, and women who dominate when most of life is indoors. Yang meant the south side of the mountain, summer, warmth, and men who dominate during times of outdoor labor. But the calendar is always changing, moving through the stages of winter into spring, summer, autumn, and winter again. Each day brings a new shifting of the proportions of yin and yang. Yin-yang changes can be visualized as wave patterns. Waves vary in amplitude and frequency and they limit what each other can do. Like a piece of music the various amplitudes and frequencies of sound harmonize with one another, or when they do not harmonize discord happens. The body contains thousands of yin-yang harmonic transformations, often working together and enabling one another during times of health; but in sick conditions, the various harmonies do not harmonize with one another and medicines need to be introduced to heat up or cool down some organ system. Chinese traditional medicine is based on the arts or rebalancing yin-yang transformation that get out of harmony. The cosmos of qi is a vast maelstrom of yin-yang transformations going into and out of harmony. This is explicitly brought to attention in taijiquan, the exercise and martial art that consists in intertwining different yin-yang patterns of qi energy.7 Nature, society, and personal and interpersonal life depend on the changes of yin and yang holding to patterns of structure. These patterns include slow-changing mountain patterns, stability of natural and social

Indeterminacy in Chinese Thought  345 environments so that human life can go on, stability of the human body so that it can do the things it must do in the day’s round, and stability of culture and thought so that people can lead intentional, communicative, and artful lives. So the interweaving of mutually supportive and limiting yin-yang movements must have some pattern. This is the problem that Plato meant to solve by saying that concrete things are always in a state of becoming, not being, passing from one form to another. Whitehead agreed with Plato and the Chinese that concrete reality is in constant flux, and said that static forms “ingress” into the flux to give it temporary passing shape.8 In Chinese thought, the earliest attempt to understand structure was with the hexagrams of the Yijing.9 In the hexagrams, six lines symbolize phases of yin or yang through which a movement passes. Broken lines are yin and solid lines are yang. The sixty-four hexagrams represent all possible combinations of yin and yang phases. Used for many purposes, from divination to esoteric numerology, the sixty-four hexagrams give an exhaustive set of descriptions of possible movements. But of course many movements have fewer or more than six phases. Moreover, alongside, within, and inclusive of any given movement are other movements. Therefore, the organic connections of movements within movements are very difficult to identify. In application to human affairs, for instance conducting a war, raising a crop, or dealing with a recalcitrant child, some hexagraphic pattern is unfolding, the tradition supposes. Suppose you have gone through a yielding yin phase, an assertive yang phase, and are now in another yin phase. Perhaps it is possible to decide to act with another yin phase of retreat, or a yang phase of advance. Whichever you choose, as king, farmer, or parent, will determine which the fourth line or phase will be, for disaster or success. Sometimes, the inertial forces of the pattern in play at first prevent any choice. But, at other times, you can make a difference to which hexagraphic pattern will emerge. In thinking about your situation in hexagraphic terms, the earlier phases might be quite indeterminate with regard to the later phases of the emerging pattern. Not only is there room for choice, sometimes, there is also room for accident as when your enemy in battle finds a sudden new ally, an unusual drought occurs, or your child is suddenly disabled. Yin-yang thinking in the Yijing is not really very practicable for science or guiding affairs, except that it calls attention to all the ways in which yin and yang movements can interrelate and affect one another. Another model for the grounds for order amid change was the notion of Heaven that flourished by the time of Confucius and Laozi; Heaven was also discussed in the Yijing, beginning with the very first hexagram. Much earlier, the Chinese had believed in a high god, Shangdi, who was a storm god similar to Yahweh or Zeus in the West. But by around 1000 bce, the literate elements of Chinese culture subordinated the personal characteristics of such a god to impersonal Heaven. Heaven was associated with the night sky, with fixed and moving stars, an impersonal but very awesome of regular and irregular

346  Robert Cummings Neville movement. As the notion of Heaven developed in diverse ways, the Chinese conceived it as the source of pattern within the cosmic changes of qi. Different interpretations of how Heaven is the source of patterns developed and by the time of the rise of Neo-Confucianism in the 10th and 11th centuries Heaven was conceived as li, Principle, or Coherence. Under the slogan, “Li is one, its manifestations many,” li was generally thought not to be a superpattern itself but rather the principle of coherence that made various pluralities of processes cohere in a pattern.10 A given plurality of changes could be made to cohere with the pattern of a mountain, another plurality with a coherent social structure, another with the patterns of an individual’s life set in a family, neighborhood, topological setting, and country. However much a given patterned harmony was changing, with its components themselves changing and its environment changing, to the extent the changes held to patter, then harmony had li, a given local manifestation of coherence. The result was a conception of the cosmos as consisting of layers upon layers, intertwinings upon intertwinings, of coherences as manifestations of li. These changes exhibiting li are all yin-yang changes with their ecological complexities. A fundamental sensibility from early to late in Chinese thought was the recognition that a lot of things just do not cohere. Not everything can be made coherent. The feistiness of qi does not take pattern or coherence easily. The reason for this is that yang can increase until it reaches its limit, but that limit might be wholly beyond the bounds of previously prevailing harmonies. Devastating floods can wash out an entire economy; barbarian hordes can overwhelm and destroy a society; a cancer can kill a body that no amount of counterbalancing can help. Similarly, on the yin side, a biological ecosystem can lose its integrity and collapse into constituent parts that cannot sustain themselves in isolation; a government can attempt to mollify the barbarians and collapse completely; the integrating vital forces in a person can retreat to death. Medical models can make it seem that the various movements of yang and yin in a person’s life are in balance, and that advance in yang is balanced by retreat in yin. But it does not work that way; no balanced harmony exists. This is precisely why medical procedures are needed: to push back against the unbalancing forces of nature. The Daodejing says that “Heaven and Earth are not humane (jen). They regard all things as straw dogs,” referring to ritual dogs made of straw that are thrown into the fire after the ceremony.11 Daoists and Confucians have stereotypically different responses to the forces of nature, society, and human life that threaten the more-or-less balanced ecologies, economies, and social and psychological organizations needed for human habitations. The Daoists identify with the forces that bend and duck, the tree that blows flat in the wind rather than the sturdy oak that is cracked in half, the water that flows down and around rather than give in to the attempts to contain it. The Daoists see nature as fearsome and seek elixirs to transcend it, spiritual practices to leave or transform the body, hermitages that avoid the big battles.12 The Confucians attempt to

Indeterminacy in Chinese Thought  347 build resources to protect the parameters of human habitation, dikes to control flooding, granaries to tide over years of bad crops, armies to oppose the barbarians, social structures built upon rituals to sustain stability, and flexibility through change. The Confucians especially see the education of sagelike qualities to be able to identify the underlying patterns of coherence and incoherence, to shore up the desirable coherences, to build new structures of coherence that unite things otherwise at odds, and to seek out the coherent structures that themselves contribute to making other desirable coherences impossible, for instance flourishing cancers.13 Daoists and Confucians agree that the human condition is fragile in the extreme, with the coherent conditions for human life under constant threat, and with a constant need for people to do something about that. I need to introduce another aspect of the cultural concept of Heaven here. Just as Heaven is that which makes determinate things coherent with their patterns of changes, so Heaven makes the heart of human nature to be an aesthetic capacity to appreciate the coherence of things and to respond in an aesthetically appropriate way. The ancient classic The Doctrine of the Mean (Zhongyong), begins “What Heaven (T’ien, Nature) imparts to man is called human nature. To follow our nature is called the Way (Tao). Cultivating the Way is called education.”14 The Confucian conception of the self is radically different from all the models in the West. A person is a continuum of elements reaching from the center of human nature to the world, “the ten thousand things.” Among the elements are the person’s body and physical environment in its various interactions, such as metabolism and growth through time; social elements, such as elementary rituals including language as well as institutions such as families, neighborhoods, economic relations and the like; and the personal educational processes. Without any of the mediating elements that differentiate people, human nature is the same in each person, the capacity to appreciate and respond aesthetically according to what coheres but with nothing to appreciate or respond to. The world of the ten thousand things is public and available to anyone, but only according to the individuating ways of their different bodies and physical environments, their different semiotically defined social elements, and their different developments in education. The goal of a sage is to be able to see the things of the world without distortion so as to grasp their coherence or value, and to be able to respond appropriately all the way through. Under ordinary circumstance, however, people’s bodies need development and their physical environments need to be arranged so that they can perceive accurately, grasp the realities of their semiotic and social environments, and perfect their skills at acting effectively. Most of the time, they are deceived by selfish physical and social structures and are incompetent actors. Education is super-important for Confucians so that the continuum between the things in the world and each person’s Heaven-bestowed nature is transparent, which the Chinese call “sincerity” (cheng). When the continuum of human embeddedness in the world is sincere or transparent,

348  Robert Cummings Neville people can appreciate the true coherences or goods of things and respond effectively and appropriately to the things according to their worths. Confucians differ significantly in their interpretations of what causes the blockages in personal engagement of the world and what to do about it in matters of education. But they conceive the world to be filled with things of value, in constant change, that people usually misperceive and with respect to which they respond with biased and incompetent ways. Human virtue is to follow the Dao by becoming educated to perceive and respond to things more appropriately, especially noting what is and is not coherent and knowing what to do about that. To summarize this already abbreviated account of indeterminacy in natural process and what human beings can do about it, the Chinese philosophical tradition (and the cultures it has influenced over two millennia) sees nature as constant changes consisting in movements of yang and yin, all modulations of qi. These changes clump into larger patterns of coherence that make up the things in the world to which we relate, always under conditions of change. The world of qi is a vast maelstrom of changes, some of which cohere with li, some of which do not. Qi does not take on coherent pattern easily, and it loses it all too easily. The human place in the world is to struggle with the forces of change that threaten the human habitat. But human beings have an inbuilt aesthetic capacity to recognize coherence where it exists and incoherence where it does not exist and might be valuable for life. Human beings also have an inbuilt aesthetic capacity to make appropriate responses to the values of things in the world, although that capacity must be developed through education. Determinateness in the world, to return to the Western question, thus consists in the patterns that happen to hold sway in the maelstrom of changes. Some of these are well reinforced by the patterns of change around them and others are undermined by both internal and environing changes. Much of nature is indeterminate, in the Chinese view, save to the expected coming of change. The opportunities for human intervention to determine things for the better are ubiquitous, although human powers for intervention might not be strong enough to make a difference for human life. The Daoists tend to look to small, immediate opportunities for spontaneous intervention, whereas the Confucians tend to look to opportunities to build human institutions that might have the capacities to contain the forces of nature and transform them into high civilization. The Daoists want to “go with the flow” but are ready to duck when the flow will kill. The Confucians want to reinforce those patterns that make for a flexible and stable human habitat, being constantly vigilant, thinking about deep structures of coherence and incoherence, looking ahead, and reconciled with a tragic sense that sometimes nothing can be done to protect or advance what is humanly good. With this sensibility, Chinese philosophers have had little occasion to imagine that the course of things is determined completely or partially by the unfolding of actuality or by rigid laws of nature.

Indeterminacy in Chinese Thought  349

Ontological Indeterminacy Ontological indeterminacy is the view that the determinate world, perhaps including some indeterminacy within it, is grounded in something that is not determinate. On the Western side, this might be a Neo-Platonic One, or a Thomistic Act of To Be, or a Cartesian creator-god who creates positive, new things. On the Chinese side, the dominant tradition has been to ignore the question of how or why there is anything at all, the ontological question. The interest in Chinese traditions, Daoist, Confucian, and Buddhist, has been mainly on what to do in our contexts of change.15 Nevertheless, a tradition has existed in China since the Daodejing that does hold to the claim that the world of process arises from something indeterminate. I shall explore this tradition shortly. First, however, let me observe that a special reason exists for interest in the ontological question in the West. Most Western symbols for the ground of being develop from conceptions of persons as agents. To the extent people can know the intentions of the gods, or of God, they can know a lot about what to expect in the world and what is valuable. To be sure, many levels of transcendence in conceptions of God exist.16 In the more anthropomorphic ones, God can be conceived as an agent acting in history on the side of his or her favorite, as in the preference of Yahweh for the Israelites over the Canaanites or the preference of Zeus for the Greeks over the Trojans. More-transcendent images of God as creator might attribute benevolence or rationality to God, so that the created world would be expected to be fundamentally good or rational. The goodness of God as a person gives rise to puzzles about theodicy, and the rationality of God has been seen as the ground of the scientific faith in the ultimate knowability of nature. To be sure, Neo-Platonic, Thomistic, and Cartesian conceptions of God, strictly speaking, do not allow for rational complexity or ordered intention within a divine mind; none of those allows for divine potentiality that might create something for a reason. But theologians and other religious people have associated those highly transcendent conceptions of God as beyond any determinateness with biblical images of God as a good, just Creator whose mind can be known through the divine Word. Thomas Aquinas had an elaborate theory of analogy that allowed him to make those associations. He claimed that goodness and rationality are positive virtues, and their opposites are corrupted with non-being. The infinite act in God must be infinite in goodness, rationality, and any of the other positive virtues, such as love, even though within the divine nature they were not determinately different from one another. If the world is created by a good and rational Creator, it would be good to know that. If the world is created by a biased and irrational Creator, it would be even better to know that. Hence, Western thought has had a strong motivation to be interested in the agency, mind, and intent of the ground of the world’s being. Chinese thought has very little if any commitment to understanding the ground of being on the analogy of a person. Rather, the Chinese metaphors

350  Robert Cummings Neville for the ground of being are like spontaneous emergence—a spring of water bursting from the ground or buds opening as winter turns to spring. I will analyze cases shortly. South Asian metaphors for the ground of being, like the West Asian ones, begin with the notion of the person. But whereas the West emphasizes the person as an intentional agent with determinate ideas and plans, South Asia thinks those are precisely the traits of persons that cause trouble—attachment, clinging, desires to possess the fruits of actions rather than only the actions themselves, as the Bhagavad Gita says. South Asian metaphors press personhood to mere consciousness, purifying the mind of reasoning and intentions, as in the emptiness of the Buddha-mind or Nirguna Brahman. The intent in South Asian religions is to identify with the indeterminate purity of the ground, or with its indeterminate emptiness. The thesis that indeterminacy gives rise to the determinate world has not been the dominant tradition in Chinese thought, which has mainly ignored the “ontological question.” I will give attention to three cases, however, in which that thesis has been expressed, from the Daodejing, from Wang Bi, and from Zhou Dunyi. The Daodejing begins (in the order of Wang Bi’s edition): The Tao (Way) that can be told of is not the eternal Tao; The name that can be named is not the eternal name. The nameless is the origin of heaven and Earth; The Named is the mother of all things. Therefore, let there always be non-being so we may see their subtlety, And let there always be being so we may see their outcome. The two are the same, But after they are produced, they have different names. The both may be called deep and profound. Deeper and more profound, The door of all subtleties.17 A straightforward reading of this is that the Dao that can be named or told of is the Dao of natural process with temporal change (including human beings and their social institutions). It is the “mother of all things” in the sense that current happenings flow from the antecedent movements of the Dao. This Dao is nameable in the sense that it has determinate features that can be described, as most of the rest of the Daodejing proceeds to do. The eternal Dao, by contrast, is the origin of the nameable Dao and it itself is indeterminate in the sense that it cannot be named. Anything that can be named, because determinate in some sense, requires being originated by the eternal Dao. The eternal Dao can be indicated (named?—surely an irony exists here) only by its function as the origin of the Dao of process. On the one hand, the eternal Dao is non-being in the sense that it has no character of its own save that it originates the process Dao. On the other hand, the eternal Dao

Indeterminacy in Chinese Thought  351 is being in the sense that it is the origin of the determinate Dao of process. The Dao of process is not separate from the eternal Dao but is simply the terminus of the originating act. If the eternal Dao did not produce anything, it would be nothing, non-being. But “after they are produced,” the originating and terminating Daos can be distinguished, despite the fact that “the two are the same.” Another way to put this is that an eternal ontological fecundity exists that grounds any movement of processive or cosmological antecedent fecundity. The eternal Dao is an ontological ground of the cosmological Dao at every moment, but is not antecedent to any moment. The being of the Dao includes its ontological arising from non-being. The eternal ontological indeterminate Dao is not a separate being or agent transcendent of the cosmological Dao but rather is the ontological dependence of the cosmological Dao on its indeterminate, nameless origin. This passage can be read in many ways, of course, and it can be assimilated to the dominant Chinese tradition that says everything is temporal and that nothing is eternal. But Wing-tsit Chan’s translation is moderate and middle-of-the-road. It is a plausible interpretive translation at the very least.18 Wangbi (226–249 ce) was a contemporary of Plotinus, living only twentythree years in the middle of the Western philosopher’s life. He was the brilliant originator of some of the most important ideas of later Confucianism and Neo-Confucianism but is most famous for his commentaries on the Daodejing and Yijing. His commentary on the passage above from the Daodejing is: All being originated from non-being. The time before physical forms and names appeared was the beginning of the myriad things. After forms and names appear, Tao (the Way) develops them, nourishes them, and places them in peace and order; that is, becomes their Mother. This means that Tao produces and completes things with the formless and nameless. Thus, they are produced and completed but do not know why. Indeed, it is the mystery of mysteries.19 It would seem that Wangbi claims there is or was a time before a first temporal creation, and his work can be read this way. But for him, the cosmological Dao that is the mother of all things is subsequent to their origin in some sense of subsequent, “develops them, nourishes them, and places them in peace and order.” This is to say, the temporal ordering and mutual relating of determinate things is caused by the cosmological Dao subsequent, in some sense, to their origin. This is compatible with the interpretation I gave of the Daodejing. Wangbi reinforces my interpretation when he writes, “All beings in the world come from being, and the origin of being is based on non-being. In order to have being in total, it is necessary to return to non-being.”20 This is Wangbi’s commentary on chapter 40 of the Daodejing: “Reversion is the

352  Robert Cummings Neville action of Tao. Weakness is the function of Tao. All things in the world come from being. And being comes from non-being.”21 It is possible to interpret this to mean that first in time there was non-being and then being with its myriad things arises. This would contrast with my interpretation that the eternal Dao is not in time, not before determinate things and their changes but eternally related to it, their eternal depth dimension. But my interpretation is straightforward and has been an undercurrent tradition in Chinese thought. Zhou Dunyi (1017–1073 ce) addresses this ambiguity directly. Zhou was one of the most important founders of Neo-Confucianism and brought the influence of Daoist and Buddhist metaphysics into that movement. In his “Explanation of the Diagram of the Great Ultimate,” he began: The Ultimate of Non-being and also the Great Ultimate (T’ai-chi)! The Great Ultimate through movement generates yang. When its activity reaches its limit, it becomes tranquil. Through tranquility the Great Ultimate generates yin. When tranquility reaches its limit, activity begins again. So movement and tranquility reaches its limit, activity begins again. So movement and tranquility alternate and become the root of each other giving rise to the distinction of yin and yang, and the two modes are thus established. By the transformation of yang and its union with yin, the Five Agents of Water, Fire, Wood, Metal, and Earth arise. When these five material forces (ch’i) are distributed in harmonious order, the four seasons run their course. The five Agents constitute one system of yin and yang, and yin and yang constitute one Great Ultimate. The Great Ultimate is fundamentally the Non-ultimate. The Five Agents arise, each with its specific nature.22 Without the alternation of yang and yin, there is no temporal succession, no time. The originating of yang and yin from the Great Ultimate is not temporal; it is their ground in every one of their temporal manifestations. In the sequence of ontological progressions Zhou traces, there is not a temporal sequence but rather an ontological sequence present in all times or changes. Real temporal flow requires everything up through the Five Agents. Controversy has surrounded the Chinese word translated “and also” in the first line of the quotation. It can also mean “in turn” and “and then.” If it is the last, it suggests that the Ultimate of Non-being and the Great Ultimate are two things. This would give a transcendent existence to the Ultimate of Non-being. I suspect myself that the Ultimate of Non-being and the Great Ultimate are two ways of looking at the same thing, as in the Daodejing, but with an ontological order. The whole sequence of things mentioned in Zhou’s statement builds up to a statement about the materially changing world, with each step a complication of what came before. Without anything having form, there would be nothing, the Ultimate of Non-being. But once the progression to complex changing things begins,

Indeterminacy in Chinese Thought  353 the Ultimate of Non-being is seen as the Great Ultimate from which yang and yin arise. The ontological creative act is regarded as the Great Ultimate insofar as it is fecund with the cosmological Dao. Considered as if nothing came from it, the Great Ultimate is the Ultimate of Non-being. Underlying the entire cosmological Dao of changes is this ontological structure of completely indeterminate Ultimate of Non-being giving rise in its very act to the Great Ultimate from which yang and yin arise, finally giving rise to the complicated determinate structures of the temporal Five Agents. The claim in this line of Chinese thought is that what explains is simpler than what it explains. The complex or ordered is what needs explanation. The only thing that does not need explanation is that which has no nature to be explained, that is, the Ultimate of Non-being. This bespeaks an astonishing (to Western cultured people) appreciation of creativity in the sense of making novel things. To be sure, there is a recognition of spontaneity in the present in which something is added to the past to get the next phase of actualization, a horizontal kind of creativity as it were, as I explained concerning cosmological indeterminacy. More profoundly, however, is the appreciation of the fact that the material rushing complexity of the present movements of things rests on existential spurts of creativity through which the world emerges from nothing, vertically, as it were. What things are is determined by the mothering of the cosmological Dao. However, that they are is built on the ontological productivity of the indeterminate, layer by layer, down to the aboriginal ontological creative act. An epistemological side of this point about the simple giving rise to the complex, the indeterminate to the determinate, has to do with intelligibility. What is ultimately intelligible? A set of first principles in terms of which things can be understood? Or the existential acts by which complex things come to be? I think the latter is surely the case for the Chinese ontological tradition I have described, and that it is true. In the long run, in the depths, to understand something is to have a kind of aesthetic perception of it being made. Because most things are temporal, intrinsically temporal, time itself is being made in the things whose existence we come to perceive with understanding. The Chinese are right, I think, to say there is something “subtle and mysterious” about the being of things in the Dao and that the grasp of this is more intellectually satisfying than seeing how the principles of things can be used to describe their order.

Conclusion The language of indeterminacy and determinateness arises out of the history of Western thinking. I hope to have shown here that similar issues are native to the history of Chinese thinking and I have expressed the issues in the thematic ways of Chinese thought. The places for indeterminate openings for human intervention and for nature’s own spontaneity are thematically dominant in China. Here is the place for cosmological indeterminacy. The

354  Robert Cummings Neville recognition of ontological indeterminacy and its spontaneous creativity is a subtheme in Chinese thought, but surely present. In our present situation, we should not think as Western philosophers or Chinese philosophers but as global philosophers. The erudition for global philosophy must include comparative philosophy. The stance of global philosophy, however, is to be erudite in all the philosophical traditions so as to craft our ideas with wisdom and perceptions from all, just as Western philosophers have learned to think with Greek, Roman, medieval, French, German, Iberian, and English ideas. This is not to say that all philosophical traditions say the same thing. Rather, they say different things about the same things.

Notes 1 On this complicated and important point, see Brumbaugh (1961). 2 On this unusual point from Descartes, see Neville (2009, chapter 13). 3 On this general problem, see Wippel (2011); in particular, May Sim’s article in that volume, “The Question of Being, Non-Being, and ‘Creation ex nihilo’ in Chinese Philosophy,” is the only article to deal with the Chinese side, and much of it is an examination of an early version of my own theory that will be expressed in the third section of the present paper. My contribution to the Wippel volume, “Some Contemporary Theories of Divine Creation,” elaborates the distinction between fullness of being and positive creation theories within the more general category of ground of being theories. 4 This trait has been present since the earliest antecedents of the Yijing. My generalizations here can be rendered more nuanced and specific by looking at some recent scholarship. Li and Perkins (2015) contains excellent essays on Chinese cosmology by contemporary scholars, including treatments of Mohism and Chinese Buddhism that I neglect in this paper. See especially the essays by Robin Wang, JeeLoo Liu, Franklin Perkins, Roger Ames, and Brook Ziporyn. See also Ames (2011, especially chapter 2). The best collection of Chinese philosophical texts in English translation is still Chan (1963); there exist now newer versions of English translations but Chan’s are fair and scholarly. In this chapter, I use the Pinyin system for spelling Chinese words, which has become the scholarly norm. But some others, including Chan, use other systems; Chan writes the Pinyin Yijing as the Wade-Giles system I Ching; he writes Dao as Tao. When I quote a translation, I use the system in the translation. For the purpose of this paper, where I assume I am writing for Western philosophers not necessarily knowledgeable about Chinese philosophy, I stick to the texts and translations in Chan’s volume. A useful table correlating Pinyin and Wade-Giles is found in Berthrong (1998), which is also a good introduction to Chinese philosophy, not only Confucianism. Berthrong (2008) is especially good on the general point that Chinese philosophy is based on change or process, not on substance. 5 Wang Yang-ming, a late 15th-early 16th-century Neo-Confucian, is famous for emphasizing the “unity of knowledge and action. See Frisina (2002) and Cua (1982). 6 See Chan (1963, chapter 11) on the Yin Yang School. The concept is pervasive in all Chinese philosophical schools, however. See the long quotation in the next section of this paper from Zhou Dunyi for a classic statement. 7 See Delza (1985). Sophia Delza was my taijiquan teacher and my foreword to her book develops the metaphor of vibratory wave patterns on yin and yang. 8 See Whitehead (1925) for his theory of ingression.

Indeterminacy in Chinese Thought  355 9 See Shaughnessy’s superb translation and commentary on the recently discovered Mawangdui text of the Yiing, the earliest text we have: Shaughnessy (1996). A beautiful coffee-table volume that includes photographs by Hugh Wilkerson is Kirk and Feng(1970). 10 See Chan (1963, chapter 32), on Cheng Yi. Principle is the source not only of pattern but of value, as I shall argue shortly. I have played upon the slogan in the title of my The Good Is One, Its Manifestations Many: Confucian Essays in Metaphysics, Morals, Rituals, Institutions, and Genders. Neville (2016, see especially chapter 1). 11 Daodejing, chapter 5, in Chan (1963, p. 141). See also Perkins (2014) and Neville (2016, chapter 15). 12 See Miller (2003). 13 See Angle (2009) both for the workings of the sage and for an explanation of the complexities of Li (Principle, or coherence). See also Angle and Tiwald (2017). 14 Chan (1963, p. 98). 15 Ames (2011) is a good expression of this understanding. 16 I have examined this in Neville (2013, especially part 2). 17 Chan (1963, p. 139). On the text of the Daodejing and the person of Laozi, including his deification, see Kohn and LaFargue (1998). 18 Translations of the Daodejing vary with the philosophy of the translator. Here is the way Philip J. Ivanhoe translates the first stanza: A Way that can be followed is not a constant Way. A name that can be named is not a constant name. Nameless, it is the beginning of Heaven and Earth; Named, it is the mother of the myriad creatures. And so, Always eliminate desires in order to observe its mysteries; Always have desires in order to observe its manifestations. These two come forth in unity but diverge in name. Their unity is known as an enigma. Within this enigma is yet a deeper enigma. The gate of all mysteries! This is from Ivanhoe (2002, p. 1). Compare this with the translation of the same text by Ames and Hall (2003, p. 77): Way-making (dao) that can be put into words is not really way-making. And naming (ming) that can assign fixed reference to things is not really naming. The nameless (wuming) is the fetal beginnings of everything that is happening (wanwu), While that which is named is their mother. Thus, to be really objectless in one’s desires (wuyu) is how one observes the mysteries of all things, While really having desires is how one observes their boundaries. These two—the nameless and what is named—emerge from the same source yet are referred to differently. Together they are called obscure. The obscurest of the obscure, They are the swinging gateway of the manifold mysteries. Ames and Hall give a strictly temporalistic, anti-eternalistic, interpretation of the texts. 9 Chan (1963, p. 321). 1 20 Ibid., p. 323. 21 Ibid., p. 160. 22 Ibid., p. 463.

356  Robert Cummings Neville

Works Cited Ames, Roger T. 2011. Confucian Role Ethics: A Vocabulary. Honolulu, HI: University of Hawaii Press. Ames, Roger T. and David L. Hall, trans. 2003. Daodejing:“Making this Life Significant”: A Philosophical Translation. New York, NY: Ballantine Books. Angle, Stephen C. and Justin Tiwald. 2017. Neo-Confucianism: A Philosophical Introduction. Malden, MA: Polity Press. Angle, Stephen C. 2009. Sagehood: The Contemporary Significance of Neo-Confucian Philosophy. Oxford: Oxford University Press. Berthrong, John H. 1998. Transformations of the Confucian Way. Boulder, CO: Westview Press. ———. 2008. Expanding Process: Exploring Philosophical and Theological Transformations in China and the West. Albany, NY: SUNY Press. Brumbaugh, Robert S. 1961. Plato on the One: The Hypotheses in the Parmenides. New Haven, CT: Yale University Press. Chan, Wing-tsit. 1963. A Source Book in Chinese Philosophy. Princeton: Princeton University Press. Cua, Antonio. 1982. The Unity of Knowledge and Action: A Study in Wang Yangming’s Moral Psychology. Honolulu, HI: University of Hawaii Press. Delza, Sophia. 1985. T’ai-Chi Ch’uan: Body and Mind in Harmony: The Integration of Meaning and Method, Revised edition. Albany, NY: SUNY Press. Frisina, Warren G. 2002. The Unity of Knowledge and Action: Toward a Nonrepresentational Theory of Knowledge. Albany, NY: SUNY Press. Ivanhoe, Philip J. trans. 2002. The Daodejing of Laozi. New York, NY: Seven Bridges Press. Kirk, Jerome and Gia-Fu Feng, trans. 1970. Tai Chi—-A Way of Centering and I Ching, Photographs by Hugh Wilkerson. New York, NY: Collier-Macmillan. Kohn, Livia and Michael LaFargue, eds. 1998. Lao-tzu and the Tao-te-ching. Albany, NY: SUNY Press. Li, Chenyang and Franklin Perkins, eds. 2015. Chinese Metaphysics and Its Problems. Cambridge, UK: Cambridge University Press. Miller, James. 2003. Daoism: A Short Introduction. Oxford: One World Publications. Neville, Robert S. 2009. Realism in Religion: A Pragmatist’s Perspective. Albany, NY: SUNY Press. ———. 2016. The Good Is One, Its Manifestations Many: Confucian Essays in Metaphysics, Morals, Rituals, Institutions, and Genders. Albany, NY: SUNY Press. ———. 2013. Ultimates: Philosophical Theology Volume One. Albany, NY: SUNY Press. Perkins, Franklin. 2014. Heaven and Earth Are Not Humane: The Problem of Evil in Classical Chinese Philosophy. Bloomington and Indianapolis: Indiana University Press. Shaughnessy, Edward L. trans. 1996. I Ching: The Classic of Changes. New York, NY: Ballantine Books. Whitehead, Alfred North. 1925. Science and the Modern World. New York, NY: Palgrave Macmillan. Wippel, John F. ed. 2011. The Ultimate Why Question: Why Is There Anything at All Rather than Nothing Whatsoever. Washington, DC: The Catholic University of America Press.

18 Cosmological Questions Ricki Bliss and Filippo Casati

For very many of us, our orientation toward the world is characterised, at least in part, by a sense of wonder: that brains work in the way they do, that gold gets spat out of dying stars, the sheer scale of the universe and the time frames involved in its creation are all features of the world we live in: features that many of us find extraordinary. Various disciplines— neuroscience, astronomy, cosmology—have developed in service to our curiosity, with their aim being to facilitate us in making sense of the world of our experience. Having answered questions about why brains behave the way that they do and how dying stars produce gold, however, one might think that we are still left to wonder why or how it is that the whole lot of it is in the first place. To wonder in this way is to be posing, and no doubt hoping for an answer to, some variation or other on a cosmological question. Cosmological questions are most typically associated, in the Western tradition at least, with cosmological arguments to the existence of God. These arguments move from purported facts about the world—say, that there is a collection of all contingent things that stands in need of explanation—to the existence of a necessary being, namely, God. Aquinas (1947), Hume (1980), Kant (1998), Leibniz (1991) and more recently van Inwagen (1996), Pruss (2006) and Rowe (1975), amongst countless others, have had much to say about the nature of these arguments, and the legitimacy of the question that have motivated them. But it is not only Western theists (and their foes) who have had a lot to say about cosmological questions. Although the background metaphysical assumptions may vary greatly, the projects of thinkers as divergent as Martin Heidegger (1996), Kitaro Nishida (2011) and contemporary metaphysical foundationalists such as Jonathan Schaffer (2010) are very much tied up in cosmological concerns.1 Although none of these philosophers is in the business of establishing the existence of God, they are very much of the view that there is a unique and suitably large collection of things whose existence stands in need of a particular kind of ultimate explanation. Broadly, then, we can think of a kind of metaphysical project that is motivated by cosmological-style questions. And it is with these questions that this short essay is concerned.

358  Ricki Bliss and Filippo Casati Attempting to answer these questions, though, is surely only a worthwhile venture if the questions themselves are good—if they are well formed and well posed. Questions such as why is there something rather than nothing, or how does anything have being whatsoever certainly seem like good questions. After all, they are perfectly grammatical, sensical and, indeed, arguably very compelling. Do we have reasons to think they are not well formed and well posed though? Plenty of philosophers have thought so. In what follows, we look at some of the reasons philosophers have offered for thinking the questions themselves are troubled. In particular, we focus on the thought that the questions are troubled because it is indeterminate what they are even asking, and we offer a discussion of how one might circumvent this concern.

Cosmological-style questions Let us begin by considering the cosmological question that is perhaps most familiar to us: why is there something rather than nothing? In the Western tradition, this question is most commonly associated with cosmological arguments for the existence of God. Such an argument can be formulated as follows: 1 There are contingent things rather than nothing at all. 2 The fact that there are contingent things has an explanation. 3 No contingent thing can explain why there are contingent things rather than nothing at all. 4 That which explains why there are contingent things must be necessary. 5 There is a necessary being that explains why there are contingent things rather than nothing at all. 6 That necessary being is God. 7 Therefore, what explains there being contingent things rather than nothing at all is God. There is much about this argument that, itself, stands in need of further explanation, but let us briefly mention two things. First, the restriction to a totality of contingent things has been justified on the grounds that no strictly (logically) necessary thing stands in need of explanation. If we assume that necessary beings are beings that simply could not fail to exist, then their being does not require explanation. On the other hand, as a contingent being could well fail to exist, its existence, it might be thought, does require an explanation.2 Second, our conclusion, as it stands, only motivates the thought that what explains the contingent collection is a necessary being. An additional argument—which we do not offer—is required to get us to the existence of God. For the sake of expediency, let us assume that the argument works. What our argument shows us, then, is that there is a collection

Cosmological Questions  359 of things whose existence needs to be explained, and there is, in fact, something that explains it, namely, God. According to Heidegger, on the other hand, the human being, thrown into the world, is surrounded by entities. There are stars lighting up the sky, hammers skillfully used by carpenters and mathematical theorems discovered by talented minds. Of course, stars, hammers and mathematical theorems are very different things: for instance, a hammer is something to be used for practical activities, while a mathematical theorem is an entity of purely theoretical contemplation. Nonetheless, they are all entities and, as such, they are something rather than nothing. But “why are there entities at all instead of nothing?”3 Heidegger spent his whole life trying to answer this question and, for this reason, it is impossible to properly discuss it here. However, for the purpose of the present paper, it is enough to say that, according to Heidegger, everything is something and not nothing because of Being. Being, which is the being an entity of any entity, is the reason in virtue of which all entities are. In other words, every entity is something rather than nothing because of Being. It is also important to note that, following Heidegger, even though Being is what makes an entity something rather than nothing, Being is not an entity itself. This assumption, called the ontological difference, is necessary because Heidegger seems to believe that no entities can explain why there are entities whatsoever.4 If so, since Being is the reason in virtue of which all entities are, Being itself cannot be an entity. During the same years, but on the opposite side of the world, Nishida Kitaro was wrestling with the same problem, trying to explain why there is something rather than nothing. However, contrary to Heidegger, Nishida does not appeal to Being, but to Absolute Nothingness. According to Nishida’s Logic of Places (2011), every object is an object because it is within a place. These notions can be understood by appealing to the judgments of the form A is B, where the subject A is subsumed by the predicate B. Nishida claims that the subject of such a judgment corresponds to an object, and that its predicate corresponds to a place within which the object is. For instance, in the judgment ‘the cat is black’, the cat is an object because it belongs to a place, namely being black. Moreover, it is also important to recall that a place can be an object as well, because it can be the subject of other judgments. For instance, in the judgment ‘Being black is a property’, being black, which was previously taken to be a place, is now an object itself because it belongs to another place, namely being a property. Given this structure, Nishida claims that there is a place in virtue of which all objects are something rather than nothing. This place is Absolute Nothingness (zettai mu). Moreover, as with Heidegger, Nishida thinks that the place that makes all objects something and not nothing is not an object itself. As such, Absolute Nothingness does not belong to any other place; otherwise, it would be the object of the place it belongs to.

360  Ricki Bliss and Filippo Casati Broadly, what we are interested in, then, is not only what we in the West tend to think of as the cosmological question—namely, the question that is supposed to motivate us to posit the existence of God—but what we are referring to as cosmological-style questions—the kinds of questions that push us to posit the existence of an ultimate explainer. For the sake of simplicity, let us take the following question as the paradigmatic cosmologicalstyle question: [CQ]: Why is there something rather than nothing? It is with the cogency of this question—and therewith the metaphysical projects associated with it—that what remains of this discussion is concerned. If there is something wrong with cosmological-style questions, this not only threatens a subset of theistic arguments to the existence of God, but the metaphysical projects of a number of philosophers across a variety of traditions.

Dismissing the Question The history of philosophy is littered with attempts to answer cosmologicalstyle questions. We are not so much interested in the answers, or even in the merits of the arguments that are associated with them. What we are more concerned with are the questions themselves. In particular, we are interested in the ways in which, and the reasons for which, the questions have been dismissed. There are several noteworthy treatments of cosmological-style questions, the upshot of which are that we shouldn’t be asking such questions in the first place. In this section, we canvas some of these treatments. In the following two sections, we offer a more extensive discussion of the suggestion that the questions are indeterminate. Cosmological questions have been charged with being ill posed or ill formed, for a number of reasons. The first amongst these is that in order for the questions to be legitimate, we must assume that (i) what is specified in the contrast is possible, and (ii) the specified totality that they claim stands in need of explanation does, in fact, need to be explained.5 Cosmological questions are generally formulated contrastively. The question, ‘Why are there any dependent things rather than nothing at all?’ looks to be a different question than, ‘Why are there any dependent things rather than only independent things?’ to take but two examples.6 Importantly, the legitimacy of the question would seem to require that what is presented in the contrast is possible. Consider [CQ]. In order for this question to be well posed, it must be possible that there is nothing in the first place. If it is impossible that there is nothing, then it is impossible that there fails to be something. And if it is impossible that there fails to be something, then there being something doesn’t, in fact, appear to cry out for explanation: The question is ill posed.

Cosmological Questions  361 Is it possible that there is nothing in the first place? The argument most commonly employed to demonstrate the possibility of there being nothing is a subtraction argument. These arguments involve subtracting entities— we could not exist, the Earth could not exist, and so on—with the idea being that, at some point, absolutely every thing could be, leaving us with exactly nothing. Several authors have pointed out, however, that these arguments are problematic. First, one might argue that it doesn’t follow from the fact that it is conceivable that we can subtract everything that it is possible that there is nothing. Second, it does not follow from the fact that for each contingent entity that can be subtracted there is a possible world in which everything is subtracted—a world in which there is nothing. To think that the possibility of nothing follows from a subtraction argument is to commit a fallacy of composition.7 Let us turn our attention, now, to (ii). Of course, one reason that the specified totality would be in need of explanation is if it is possible that it does not exist.8 If it is not possible that there is nothing in the first place, then we seem to have a reason to believe that there being something does not cry out for explanation.9 Are there other reasons to suppose that the specified totality may not need explaining? The assumption that the collection stands in need of explanation appears to rely on an application of some version or other of the Principle of Sufficient Reason (PSR). Much has also been said in the literature about the PSR.10 According to the full-blown PSR, every fact has an explanation. Results from quantum physics, however, would seem to indicate that this principle simply isn’t true. It seems that there are some (contingent) facts that do not have explanations. Indeed, the very idea of a brute contingent fact has gained a lot of traction in the contemporary literature. Perhaps the fact that there is a totality of contingent things is one such fact. The best known of the objections to the thought that there is a totality of contingent things that needs an explanation is surely, however, taken from Hume (1980, Part 9). To explain each contingent entity in terms of its causes, thinks David Hume, is to explain everything that stands in need of explanation. In other words, there just is nothing that remains to be explained. There simply is no further fact that needs to be explained having explained all the particular local matters of fact. Much ink has been spilled addressing aspects of the afore-mentioned problems. Although doubtless more could be added to the existing debates, we wish to focus on a reason to take issue with cosmological questions that we believe has not been discussed extensively in the literature. More recently, it has been suggested that cosmological-style questions run into another kind of problem: they are indeterminate.11 In the remaining two sections of this chapter, we take up this issue. If it is true that these questions are indeterminate, then this will have significant consequences for how cosmological-style questions can be answered. If this is not true, then we have eliminated one further reason for thinking there is a problem with the questions in the first place.

362  Ricki Bliss and Filippo Casati

Cosmological-style questions and indeterminacy Suppose we were to ask how many flamingos there are in the Rann of Kutch. Whilst we would be very surprised to discover that this question was something that your average citizen would have an answer for, we would all likely agree that someone knows the answer: no doubt there is at least one world expert whose business it is to study the movements of flamingos at various spots across the Earth through the seasons. But even if such an expert did not exist, we would surely agree that it is, in principle, possible to count how many flamingos are there. Now suppose we were to ask you how many brown things there are before you. If your location is anything like ours, there will be a number of brown things available to start counting off: a desk, a hair comb, a dead plant. Suppose now that in the middle of your counting exercise, a friend arrives and you ask her how many brown things she can see. Like us, she starts to count, only she arrives at a different number to you. When pressed, she smiles, and points out that where you had counted a desk as one item, she had seen four things: a desk with three drawers under the top. Although you’ve been a bit slow off the mark, fortunately you’ve taken a few philosophy courses, and you realise fairly quickly that the problem is worse than this. Much worse. Before you there are brown desk legs, brown leaves, brown soil, brown bits of soil, brown corners of brown tables, and so on. In fact, now that you’re going, you realise that there are, likely, infinitely many brown things before you. Whilst you are correct to note that there are many more brown things in front of you than you could ever have the time, or inclination, to count, you have likely failed to notice something significant: it is, in fact, completely indeterminate what the brown things actually are. Or, to put the point more finely, it is indeterminate what the term ‘brown thing’ is actually supposed to apply to. This, then, is what is known in the literature as the counting problem.12 The counting problem is a problem thought by many to be of both deep metaphysical and cognitive significance, and the reason for this is that it is supposed to draw our attention to the important need to distinguish between different kinds of sortal terms: substantive sortals and dummy sortals. We don’t wish to use the counting problems as a reason to believe this distinction ought to be drawn, but we do think the problem illuminates something interesting about the way language works in this connection. Substantive sortal terms are terms such as ‘flamingo’, ‘person’, ‘set’ and ‘gold’.13 Sortal terms are to be contrasted with adjectives: ‘brown’, ‘lazy’, ‘tempestuous’. Sortal terms and adjectives both have associated with them application conditions.14 The application of these terms delivers items that fall within their extension. What sortal terms have associated with them, however, that adjectives do not, are identity conditions, or what Amie Thomasson refers to as coapplication conditions,15 and conditions of

Cosmological Questions  363 individuation. Crudely put, identity conditions are what allow us to identify and then re-identify an object. Whilst one can wonder if the computer at which one writes is the same computer that was there yesterday, one cannot wonder if the tempestuous here now is the same tempestuous that was encountered last week. What we can do, though, is wonder if this tempestuous weather is the same that was experienced last week when on another continent. Sortal terms denote sortal concepts, which, in their turn, are thought to pick out kinds or categories. So the term ‘flamingo’ denotes the concept flamingo, which picks out the kind flamingo.16 Unlike substantive sortal terms (‘flamingo’, ‘gold’), dummy sortals function grammatically like (substantive) sortals, but not logically. What this means is that although we employ dummy sortals in our languages as though they denote kinds, what distinguishes a dummy from a substantial sortal is that the latter, and not the former, have associated with them criteria of identity and individuation. A criterion of identity tells us when one thing that a sortal term applies to is the same thing, numerically speaking, as something else that the term applies to. It is a criterion of identity that allows one to determine whether the pyjamas being worn now are the same ones worn last night. According to Lowe,17 a criterion of identity will be of something like the following form: ‘If x and y are Ks, the x is identical with y if and only if x is Rx-related to y’, where K is a sortal term and Rx a non-identity equivalence relation on Ks. That sortal terms have criteria of identity associated with them is the reason that we can co-apply terms such as ‘pyjama’ successfully. A principle of individuation, however, is that which allows us to identify something as being the thing that it is. In the case of sets, for example, what individuates them are their members. Humans, on the other hand, are going to be individuated (as humans) by dint of having a certain type of ancestry, or certain capacities, depending upon what one’s particular theory is. For illustrative purposes, it might help to think of what individuates us as being part of our essence (but one does not need to commit to essentialism to think that objects are individuated). In order, then, to be considered a sortal, a term must meet the following conditions: 1. Sortal terms have associated with them criteria of application and coapplication. 2. Sortal terms have associated with them determinate criteria of identity and individuation, where items that fall under a sortal term are governed by the same criteria of identity and individuation. To understand the significance of these points for our discussion of cosmological questions, let us return, first, to the counting problem. Again, although we do not believe that the counting problem provides us with a reason to believe there is a distinction between dummy and substantial sortals, we do believe that it provides us with a useful means by which we can

364  Ricki Bliss and Filippo Casati gain some insight into the distinction. Consider, then, how many contingent things there are in the room you are sitting in. You, yourself, seem like a contingent thing, as does the chair you sit at and the device you are reading this chapter on. These items should be on our list. But what about more recherché items, such as the bottom square inch of the leg of your chair and the last atom in your left big toe? Or the boundary between the last table atom on the right-hand bottom corner of the table leg and the first atom in the appropriate region of the floor? Or the lungful of air you just inhaled? Should these make it onto our list? We imagine many a reader crying, ‘Yes! All of those things should be on the list!’, and we are happy to agree with this. Note, however, that by the time we are singling out items such as the last atom in your left big toe, we are associating those items with terms that have criteria of identity and individuation (and therewith conditions of application and coapplication) associated with them. We are able to count things, regardless of how weird they might be, once we are clear on the kinds of things we ought to be counting.18 Returning to [CQ], we can note that it appears to employ two quantifier terms—‘something’ and ‘nothing’. We will focus on interpretations of ‘nothing’ in a moment, but for now we would like to focus, instead, on ‘something’. It is not necessary that we interpret ‘something’ as a quantifier term in [CQ]. Instead, we might wish to slightly modify [CQ], and understand it as ‘Why is there some thing rather than nothing?’ What we have appearing in [CQ], then, is the sortal term ‘thing’. So interpreted [CQ], and cosmological-style questions more broadly, appear to be in trouble, for the term ‘thing’ is a dummy rather than a substantive sortal. Recall from (ii) above, that it is a condition on counting as a substantive sortal term that that term has associated with it determinate criteria of identity and individuation—with items falling under the term being governed by the same criteria. In the case of ‘thing’, unfortunately, the term has no such determinate criteria associated with it. Consider the device you are reading this book on and your mother. Both are, presumably, things, and yet both seem to be governed by different criteria of identity and individuation: what it is to be a mother is very different to what it is to be a reading device. Items that fall under the term ‘thing’ are not governed by the same criteria of identity and individuation. And because the term does not have determinate criteria of identity and individuation associated with it, it is indeterminate, so we believe, what the term refers to in the first place. The problem does not stop here, however. Because it is indeterminate what the term ‘thing’ is supposed to refer to, it is also indeterminate which of the things are still in existence that were in existence five minutes ago. The reason for this is that our criteria of identity help fix what the persistence conditions on objects are. I cannot, in fact, correctly reapply the term ‘thing’ if it is indeterminate what the term was supposed to refer to in the first place: I can’t reapply a term if I can’t apply it in the first place. In other

Cosmological Questions  365 words, where ‘thing’ has no conditions of application, it does not have conditions of co-application either. In an important sense, the problem here seems to be one of generality: the term ‘thing’ is simply too general to have associated with it that which is required to allow it to function both semantically and logically like a substantive sortal term. The exact same set of problems can be generated whether we are talking about ‘blue things’, ‘hot things’ or ‘contingent things’. Indeed, we believe, the same problems will be generated for other sufficiently general terms such as ‘entity’ and ‘object’, as well.19 The suggestion that terms such as ‘thing’ are dummy sortals seems peculiar, though; after all, we very often ask questions about things, entities and so on, and seem to do a pretty good job of answering them. We know exactly what someone is asking us when they ask how many things we received for Christmas; just as we know what a philosopher is talking about when they develop theories about contingent entities. What we seem to have arrived at, then, are two seemingly inconsistent views regarding a certain subset of sortal terms. On the one hand, we have suggested that terms such as ‘thing’ are dummy sortal terms without criteria of identity and individuation associated with them: questions that employ them are defective because it is indeterminate what the terms are referring to. On the other hand, we have suggested that we successfully use terms such as ‘thing’ all the time. How can we reconcile these claims? The use of the term ‘thing’ as a substantive sortal term in cosmologicalstyle questions gives rise to defective questions. Let us call this use of ‘thing’ the defective use. According to the defective use, we employ terms such as ‘thing’ as though they are substantive sortal terms; as though there is some kind, namely thing, the instantiation of which deserves an explanation (an explanation for why any member of that kind exists whatsoever).20 But as it is indeterminate what the term ‘thing’ is supposed to refer to, questions that employ such terms do not carry content, and cannot properly be answered. In other words, the questions are defective owing to the defective use of the relevant sortal term. We can contrast the defective use of ‘thing’ with what we can call the covering term use.21 According to the covering term use, terms such as ‘thing’ act as a covering term for lists of (likely contextually constrained) substantive sortal terms. To ask why there are any things whatsoever is just to ask why there are turtles, peonies, watches and so on. We are able to sensibly ask and answer questions about things exactly because we take the term ‘thing’ to be standing in, as it were, for terms that pick out genuine kinds. It is important to note, however, that on the covering term use of the relevant sortal terms (‘thing’, ‘entity’, ‘being’) the questions look to be somewhat neutralised. What we mean by this is that cosmological-style questions become questions about fairly ordinary entities—flamingos, pianos, and so on. Moreover, so understood, it seems we have the resources to answer such

366  Ricki Bliss and Filippo Casati questions. Science in particular provides us with the means to explain why there are people, flamingos and so on.22 Where this leaves the cosmologically concerned metaphysician, however, would seem to be between a rock and a hard place. If cosmological-style questions are defective then it seems reasonable to suppose that we ought not be asking them. If, on the other hand, the questions are non-defective—because we assume the relevant sortal term to be acting as a covering term—then one might think we have at our disposal relatively straightforward ways of answering them. But if this is the case, far from needing such exotic entities as God, Absolute Nothingness or Being to answer cosmological-style questions, we need only appeal to ordinary, worldly explainers. In either case, it would appear that attempts to engage with cosmological questions, and the elaborate metaphysical projects that have been built up around them, have been largely misguided.

Cosmological-style questions and determinacy If the foregoing assessment is correct, then Heidegger, Nishida, Leibniz, and very many others have been embarked on projects founded upon a basic misunderstanding. Is there anything that can be said by way of response to the charge that cosmological style questions are either indeterminate and defective or determinate and easily answerable? We believe, in fact, that much can be said. For reasons of economy, however, we must limit our discussion. For present purposes, we choose to accept that (i) the relevant sortal terms are dummy sortal terms, (ii) on one interpretation of the relevant questions, they are indeterminate and defective, and (iii) the covering term analysis of the relevant dummy sortal terms is such that cosmological style questions can be understood as asking questions about items that fall under substantive sortal terms. In order to see how we might resist the aforementioned conclusions, however, let us now spend some time trying to understand the question a bit better. How exactly we are supposed to understand [CQ] is a difficult and contentious matter. And anyone familiar with the relevant literature will know that much has been said about the problems with asking and answering such questions. For our purposes, however, there are two features of [CQ] that we believe require immediate attention. The first of these is that it is to be understood contrastively; and the second returns us to the issue of quantifier terms. First to [CQ] as a contrastive question. On the covering term use of ‘thing’, we are encouraged to understand ‘thing’ as simply being shorthand for a long list of substantive sortal terms. To ask why there is something rather than nothing is to ask why there are pianos, people and so on. The suggestion that we can answer this question because science tells us why there are pianos, why there are people, and so on, simply looks to be mistaken. Whilst the results of science help us to explain why there are Homo

Cosmological Questions  367 Sapiens, for example, it is not at all clear that they help us to explain why there are people rather than nothing—at best we are able to explain why there are Homo Sapeins rather than not. Understanding [CQ] as asking why there are people, why there are planets, and so on—where to be able to answer each question is to be able to answer the question—seems to fail to respect the question as it is initially posed. But even where we accept the covering term use of ‘thing’, [CQ] does not appear to be asking why there are pianos rather than nothing, flamingos rather than nothing, and so on. Instead, what it seems to be asking is why there are planets, flamingos, car parts, electrons . . . rather than nothing (for a likely infinitely long list of things). Understood in this way, our alleged ability to answer [CQ] naturalistically presents us with a problem: any entity invoked in a naturalistic explanation will itself be on the list of things whose existence stands in need of explanation. The problem here is not that we may not be able to terminate our explanatory regresses, but, rather, that we appear to invoke in our explanans items appearing in the exlanandum. We are not at all convinced that the naturalistic answers to the covering term use of [CQ] are as straightforwardly acceptable as some would have us believe.23 If it is, indeed, the case that naturalistic answers to [CQ] fail, this is all the better for the likes of Heidegger and Nishida. Indeed, for both Heidegger and Nishida, the reason in virtue of which there are entities (rather than nothing) is not itself an entity. Contrary to this idea, in order to explain why there are entities, such as pianos and flamingos, naturalistic explanations appeal to other entities, such as piano keys and flamingo’s feathers. But this is exactly what Heidegger and Nishida would deny. Now to our second point. On one very straightforward interpretation of [CQ] it employs two quantifier terms—‘something’ and ‘nothing’. The argument from dummy sortals requires that we treat ‘something’ not as a quantifier term but rather as a involving a sortal term—some things. In spite of a willingness to treat ‘something’ as a noun phrase, the idea persists that ‘nothing’ needs to be interpreted as a quantifier. as an example, Peter Van Inwagen interprets the expression ‘there is nothing’ as ‘there is an empty world’, namely a world with no things inside.24 And this is the case exactly because ‘nothing’ is understood as a quantifier: ‘there is nothing’ simply means that ‘there are no things’. Now, if we assume this understanding of ‘nothing’, we can rephrase [CQ] in the following way: why are there some things rather than no things? Furthermore, we can break [CQ] into two sub-questions: [CQ1]: Why is it the case that there are some things? and [CQ2]: Why is it not the case that there are no things at all?

368  Ricki Bliss and Filippo Casati It seems evident that these two sub-questions are concerned with the same issue. Since, if it is the case that there are some things, it is not the case that there are no things (and vice versa), it is enough to answer one of the two sub-questions (either [CQ1] or [CQ2]) in order to answer [CQ]. If so, we have, at least, two available strategies to answer [CQ]. The first strategy is [S1] to show why it is the case that there are any things whatsoever. Such a strategy has been applied by many thinkers in the history of philosophy: for instance, in order to explain why there are things whatsoever, Heidegger appeals to Being25 and Nishida to Absolute Nothingness.26 The second strategy is [S2] to show why it is impossible (or, at least, highly unlikely) that there are no things at all. This second strategy has been recently explored by Van Inwagen (1996). Indeed, he argues that the fact that ‘there is nothing’, namely the fact that ‘there is an empty world’, is not impossible but as improbable as anything can be. But what happens if we challenge the idea that ‘nothing’ can be understood only as a quantifier? It is important to specify that we do not want to deny that ‘nothing’ is, most of the time, taken to be a quantifier. We are well aware that when we say that there is nothing in the fridge, we mean that the fridge contains no things: no beers, no cheese, no bread—nothing, indeed! However, in the rest of the present paper, we aim at showing that, first of all, it is possible to treat ‘nothing’ as a singular term. As singular terms cannot be dummy sortal terms, ‘nothing’ cannot be a dummy sortal term either. Secondly, we aim at showing that, using ‘nothing’ as a singular term, it is possible to engage with [CQ] without using any dummy sortal term. As such, [CQ] will not be ill formed. To begin with, looking at literature, theology and figurative art, it seems that nothing can be treated as a singular term. In a painting of Dali, a chemist desperately seeks something important, namely ‘nothing’. In Musil’s Posthumous Papers of a Living Author, a gladiator sees a bright light coming from the sky and he names it ‘nothing’ while, according to medieval theological interpretations of the Bible, God creates the world out of it. Substantial metaphysical claims are not an exception. For instance, when the medieval philosophers tried to understand why ex nihilo nihil fit—namely why nothing comes from nothing—they intuitively took the first occurrence of ‘nothing’ as a quantifier and the second occurrence as a singular term. It is easy to get confused in talking about nothing exactly because ‘nothing’ can be understood both as a quantifier and as a term. Think about the weird Tea Party attended by Alice in Wonderland. She sits around a table with the Mad Hatter and the March Hare. “ ‘Take some more tea’, the

Cosmological Questions  369 March Hare said to Alice. ‘I have nothing yet’, she replied in an offended tone: ‘so, I cannot take more.’ ‘You mean you cannot take less,’ said the Hatter. ‘It is very easy to take more than nothing.’ ”27 As mentioned above, here the confusion is generated by the fact that Alice uses ‘nothing’ as a quantifier and the Hatter uses ‘nothing’ as a singular term. Such a confusion would not have occurred if the Hatter, knowing a bit of philosophy, would have added to ‘nothing’ (used as a singular term) the suffix ‘-ness’ or the definite article ‘the’. In the rest of the paper, to avoid the confusion with the quantifier, we will write ‘nothing’ (understood as singular term) in boldface (nothing). Since we have established that it is legitimate to use ‘nothing’ as a singular term, it is also legitimate to ask what happens to [CQ] if we interpret ‘nothing’ as nothing. Well, apparently—but only apparently!—not much. Substituting ‘nothing’ with nothing, we can rephrase [CQ] in the following way: why is there something rather than nothing? Let’s call this new question [CQ*]. Furthermore, we can also rephrase the two relative sub-questions introduced above ([CQ1] and [CQ2]) into [CQ*1] and [CQ*2] obtaining: [CQ*1]: Why is it the case that there are some things? and [CQ*2]: Why is it not the case that there is nothing? At this point, the analogy between [CQ] and [CQ*] seems to become less straightforward. As we have seen, sub-question [CQ1] and sub-question [CQ2] actually ask about the same issue. And this is also the reason why, in the case of [CQ], answering either [CQ1] or [CQ2] is enough to find a satisfactory answer for [CQ] too. But is this true for [CQ*] as well? If it is the case that, when there are some things, there is not nothing (and vice versa), then, in order to answer [CQ*], it is enough to answer one of the two sub-questions (either [CQ*1] or [CQ*2]). As such, we can answer [CQ*] applying one of the two, appropriately revised, strategies discussed above. On the one hand, we can show why it is the case that there are any things whatsoever ([S*1]). On the other hand, we can show why it is impossible (or, at least, highly unlikely) the case that there is nothing ([S*2]). Now, the fact that there is nothing implies that there are no things (and vice versa) if and only if nothing is understood and interpreted in certain ways. Showing that nothing can be understood as a singular term is clearly not enough. It is easy to see why, considering the following situation. Let’s imagine that, at the bottom of the Indian Ocean, a scientist discovers a primitive organism. This primitive organism is so microscopic that she decides to name it ‘nothing’. After that, the scientist writes a book, entitled The Book of Nothing, in which she develops a theory according to which every living creature is the result of the evolution from that primitive microscopic

370  Ricki Bliss and Filippo Casati organism, called nothing: in a sense, everything comes from nothing! In this case, using the term ‘nothing’ as the scientist does, it is not true that if there are some things than there is not nothing. Indeed, nothing seems to be something itself: it is just the name of a micro-organism discovered at the bottom of the Indian Ocean! So, which understanding of nothing is compatible with the idea that, if there is something, then there is not nothing (and vice versa)? In which way should we interpret the singular term ‘nothing’? One possible way is to consider nothing to be a sort of global absence, namely the absence of everything. At the end of the day, it is intuitive to think that nothing is “absolutely nothing: it is nothing at all.”28 Following the argument by subtraction, nothing is exactly what we obtain when we remove everything from the totality of all things. Now, take the totality of everything (chairs, tables, atoms . . .) and start to subtract a first thing (a chair), a second thing (a table), a third thing (another chair), until—poof !—there is nothing, nothing at all, namely the absence of everything. Until here, we have seen that we intuitively know what we are talking about when we talk about nothing as a singular term. However, for philosophically trained minds, this may not be enough. More details are needed and, at this point, some contemporary philosophers can help us. Let’s start with Oliver and Smiley. According to them, the singular term nothing is an empty term: it does not refer to anything. Another term that behaves like ‘nothing’ is ‘zilch’. In both cases, “it is already well established [that they] indicate non-entity or nil quantity or more generally nothing.”29 When we say that “[a] ham sandwich is better than zilch”, we mean that a ham sandwich is better than no food at all: it is better than the absence of food, namely zilch itself.30 In the same way, when we say that “there is something rather than nothing,” we mean that there is something rather than the absence of ham sandwiches and chairs and table and stars and . . . everything else. According to Oliver and Smiley, in both cases, zilch and nothing are empty singular term. Now, to see the reason why Oliver and Smiley’s account of nothing can be unsatisfactory, let us appeal to their distinction between weak and strong predicates. As Casati and Fujikawa correctly summarizes: “An n-place predicate F is strong with respect to its i-th place if it is analytic that F(. . ., a,. . .) is false whenever a is empty, where. . ., a, . . . is a sequence of n terms whose i-th member is a; otherwise it is weak with respect to its i-th place. They present four kinds of weak predicates, that is, (i) weak identity ≡, (ii) nonexistence predicates, (iii) semantic predicates which deal with empty terms and weak predicates, and (iv) negation of strong predicates.” 31 This list of weak predicates doesn’t need to be an exhaustive one. However, the weak predicates discussed above are already enough to see why such an account of nothing can be problematic. For instance, according to Oliver and Smiley, ‘being different from’ is a strong predicate.32 If so, the sentence ‘nothing is different from the table’ is false and this doesn’t seem

Cosmological Questions  371 good because, in understanding why there is something rather than nothing, we may want to say that nothing is something metaphysically ‘special’: at the end of the day, we want to say that nothing is the absence of everything and that, as such, it is different from a table. This is not possible, following Oliver and Smiley. Another possibility is to accept the idea that nothing refers to something, namely to the absence of everything. This is possible assuming an ontology in which absences are included in the so-called furniture of the world. As McDaniel (2014) suggests, given the way in which we use quantificational expressions, it seems to be true that the absence of Fs exists if and only if there are no Fs. At the end of the day, we very often deal with absences: for instance, shadows are absences of light and holes are absences of matter. Moreover, if you are an ontological pluralist à la McDaniel and you believe that different things exists in different ways, then you may want to claim that absences exist in different ways to tables and chairs. In a sense, absences are less real. However, absences still exist. Now, if you endorse such a stance about absences, then nothing is just a singular term, which refers to a particular kind of absence, namely the absence of everything. The negative outcome is that, following this understanding of nothing, [CQ*] seems to be trivialized. Indeed, it seems that the reason why there is something rather than nothing is that, even when there is nothing, there is something, namely the absence of everything. In other terms, someone may worry that such an understanding of nothing makes [CQ*] uninteresting. As in the case of Oliver and Smiley, this second understanding of nothing is far from being ideal. However, we should not forget that the aim of the present discussion is not to show that interpreting nothing as a singular term makes cosmological questions interesting. The aim is simply to show that cosmological questions are not necessarily ill formed, and that one way to do so is to interpret nothing, not as a quantifier, but as a singular term. If the price that we pay in doing so is to turn the cosmological question into an uninteresting riddle, then so be it.

Conclusion Many philosophers, perhaps more than commonly realised, have had at the heart of their metaphysical projects a concern with cosmological-style questions. The disavowal of these questions—a not uncommon position in the modern Western tradition—threatens more than traditional (Western) theistic arguments for the existence of God. If cosmological-style questions are in trouble, then so are the metaphysical projects of thinkers such as Heidegger and Nishida. Some might think that the more reasons we have to disregard the projects of these thinkers the better. We, on the other hand, are of the view that the likes of Heidegger and Nishida have much to teach us about how we ought to do metaphysics. If cosmological-style questions are troubled because they are either indeterminate or determinate but easily

372  Ricki Bliss and Filippo Casati (naturalistically) answerable, much more would seem to be under threat than theistic arguments to the existence of God. We hope to have shown that even where we might agree that terms such as ‘thing’ are dummy sortals, the journey from there to the neutralization of the question is, nonetheless, marked with perils.

Notes 1 That these philosophers are engaged in cosmological-style projects is not always made clear. We believe that this is partly (perhaps even largely) because these thinkers are not concerned with establishing the existence of God. Why we believe Heidegger and Nishida to be on such projects will become clearer as this paper progresses. 2 To ask why there is something (rather than nothing) is very often interpreted, then, as a question asked of a restricted domain. 3 Heidegger (2000, p. 1). 4 See, for instance, Nicolson (1996). 5 See Grünbaum 2004 for an excellent discussion of the many problems that are believed to beset cosmological arguments. 6 See Bliss (forthcoming) for a discussion of fundamentality and contrastive questions. 7 See Grünbaum (2004, p. 564). It is fallacious to think that the properties of the parts are necessarily properties of the whole. 8 Grünbaum (2004, pp. 567–569). 9 Here is a good example of a case in which we can see that subtle variations on the question will make a difference. The question ‘why does this particular collection of things exist rather than some other?’ demands of us a very different set of considerations than the question ‘why is there something rather than nothing at all?’ 10 See Della Rocca (2010) and Pruss (2006) for excellent discussions of the PSR. 11 Maitzen (2013). 12 See Lowe (2009), Maitzen (2012, 2013), and Thomasson (2007). 13 Not everyone agrees that non-count nouns—mass nouns such as ‘gold’—ought to be considered sortal terms. We, however, follow Lowe (2009, pp. 13–14), and consider mass nouns as sortal terms, as they have criteria of identity associated with them (even though we cannot count them). 14 Thomasson (2007, pp. 39–45). 15 Ibid., pp. 40–44. 16 See Lowe (2007) for a particularly informative discussion of sortals. 17 Lowe (2007, p. 515). 18 Perhaps a slightly more informal way of putting the point is that there just is no way to be a thing that is not a flamingo, notepad, asteroid, etc. (something denoted by a substantive sortal term). See, for example, Maitzen (2012) and (2013), and Lowe (2009, pp. 14–15). 19 See Maitzen (2012). 20 See Maitzen (2013) for an excellent and sustained elaboration of this point. 21 See Thomasson (2007), especially ch.6 for the notion of a covering term. See Maitzen (2013) for a discussion of the covering use of terms in the context of cosmological arguments. 22 See Maitzen (2013) for a discussion of these points. 23 Maitzen (2012) and (2013) makes a compelling case for the covering term use of ‘thing’ in cosmological questions. He also argues that the questions are then answerable naturalistically. 24 Van Inwagen (1996).

Cosmological Questions  373 25 Heidegger (1996). 26 Nishida (2011). 27 Carroll (1865, p. 78). 28 Priest (2014, p. 151). 29 Oliver and Smiley (2013, p. 602). 30 cf. Oliver and Smiley (2014, p. 602). 31 Casati and Fujikawa (2015, p. 5). 32 Oliver and Smiley (2014, p. 603).

Works Cited Aquinas, Thomas. 1947. Summa Theologica, translated by The Fathers of The English Dominican Province, Notre Dame: Christian Classics. Bliss, Ricki. forthcoming. “What Work the Fundamental.” Erkenntnis. Carroll, Lewis. 1865. Alice’s Adventures in Wonderland. London: McMillan & Co. Casati, Filippo and Naoya Fujikawa. 2015. “Better than Zilch?” Logic and Logical Philosophy. 24: 255–264. Della Rocca, Michael. 2010. “PSR.” Philosophers’ Imprint 10(7): 1–13. Grünbaum, Adolf. 2004. “The Poverty of Theistic Cosmology.” British Journal for the Philosophy of Science. 55: 561–614. Kant, Emmanuel. 1998. Critique of Pure Reason, translated byGuyer, Paul and Allen Wood, Cambridge: Cambridge University Press. Heidegger, Martin. 1996. Being and Time, translated by Joan Stambaugh. New York: SUNY Press. Hume, David. 1980. Dialogues Concerning Natural Religion. Indianapolis, IN: Hackett. Kant, Immanuel. 1998. Critique of Pure Reason, translated by Paul Guyer and Allen Wood. Cambridge: Cambridge University Press. Leibnitz, Gottfried Wilhelm. 1991. Discourse on Metaphysics and Other Essays, translated by Daniel Garber and Roger Ariew. Indianapolis, IN: Hackett. Lowe, E. Jonathan. 2009. More Kinds of Being: A further Study of Individuation, Identity and the Logic of Sortal Terms. Oxford: Wiley-Blackwell. ———. 2007. “Sortals and the Individuation of Objects.” Mind and Language 22(5): 514–533. Maitzen, Stephen. 2013. “Questioning the Question,” in The Puzzle of Existence: Why Is There Something Rather Than Nothing? edited by Tyron Goldschmidt. New York, NY: Routledge. ———. 2012. “Stop Asking Why There’s Anything.” Erkenntniss. 77: 51–63. McDaniel, Kris. 2014. “Ontological Pluralism and the Question of Why There Is Something Rather Than Nothing”, pp. 290-320, in The Philosophy of Existence: Why There Is Something Rather Than Nothing, edited by T. Goldshmidt. New York, NY: Routledge. Nishida, Kitaro. 2011. Places and Dialectic. Oxford: Oxford University Press. Oliver, Alex and Timothy Smiley. 2013. “Zilch.” Analysis. 73: 601–613. Priest, Graham. 2014. “Much Ado About Nothing.” Australasian Journal of Logic. 11: 146–158. Pruss, Alexander. 2006. The Principle of Sufficient Reason: A Reassessment. Cambridge: Cambridge University Press. Rowe, William L. 1975. The Cosmological Argument. Princeton: Princeton University Press.

374  Ricki Bliss and Filippo Casati Schafferm, Jonathan. 2010. “Monism: The priority of the Whole.” Philosophical Review. 119(1): 31–76. Thomasson, Amie. 2007. Ordinary Objects. Oxford: Oxford University Press. Van Inwagen, Peter. 1996. “Why is there anything at all?” Proceeding of the Aristotelian Society. 70: 95–110.

Contributors

Bruce Ellis Benson is Senior Research Fellow at the Logos Institute, University of St Andrews. He is the author or editor of thirteen books. Dr. Benson serves as the Executive Director of the Society for Continental Philosophy and Theology. Ricki Bliss is Assistant Professor of Philosophy at Lehigh University, Pennsylvania. Previously, she was a Visiting Lecturer at Otago University, Dunedin, New Zealand, and a Japan Society for the Promotion of Science Research Fellow at Kyoto University, Japan. She has also been the recipient of an Alexander von Humboldt Research Fellowship, which was taken up at the University of Hamburg, Germany. Nahum Brown is a Research Fellow at Sun Yat-sen University. He is the editor (along with William Franke) of Transcendence, Immanence, and Intercultural Philosophy (Palgrave, 2016). He is also editor (along with J. Aaron Simmons) of Contemporary Debates in Negative Theology and Philosophy (Palgrave, 2017). He is currently working on a manuscript entitled Hegel on Possibility. G. Anthony Bruno is Lecturer in Philosophy at Royal Holloway University of London. Previously, he was SSHRC Postdoctoral Research Fellow at McGill University, Toronto, Faculty Lecturer at University of Toronto Scarborough, and Alexander von Humboldt Postdoctoral Research Fellow at University of Bonn. He has published numerous articles on Kant, German idealism, and phenomenology, and is the editor of Schelling’s Philosophy: Freedom, Nature, and Systematicity (Oxford University Press) and co-editor of Skepticism: Historical and Contemporary Inquiries (Routledge). Filippo Casati is a JSPS postdoctoral fellow at the University of Kyoto. He is also the recipient of a two-year research grant from the Alexander von Humboldt Foundation. He works on the intersection between analytic philosophy (dialetheism, Meinongianism), Heidegger’s fundamental ontology and Japanese philosophy. His research has been published in both international journals and edited volumes. Casati is one of the

376 Contributors editors of the second edition of Routley’s Exploring Meinong Jungle and Beyond and, with Graham Priest and Chris Mortensen, is also editing a special issue for the Australasian Journal of Logic. Steven G. Crowell is Joseph and Joanna Nazro Mullen Professor of Humanities and Professor of Philosophy at Rice University. He has authored numerous articles on phenomenology, as well as two books: Normativity and Phenomenology in Husserl and Heidegger (2013) and Husserl, Heidegger, and the Space of Meaning: Paths Toward Transcendental Phenomenology (2001). He edited The Cambridge Companion to Existentialism (2012) and, with Jeff Malpas, Transcendental Heidegger (2007). Currently he co-edits Husserl Studies with Sonja Rinofner-Kreidl. William Desmond is David Cook Chair in Philosophy at Villanova University, and Professor of Philosophy Emeritus at the Institute of Philosophy, Katholieke Universiteit Leuven. His work is primarily in metaphysics, ethics, aesthetics, and the philosophy of religion. He is the author of many books, including the groundbreaking trilogy Being and the Between (1995), Ethics and the Between (2001), and God and the Between (2008). Being and the Between was winner of both the prestigious Prix Cardinal Mercier and the J.N. Findlay Award for best book in metaphysics. Other books include Art and the Absolute (1986); Desire Dialectic and Otherness: An Essay on Origins (1987; 2nd ed. 2014); Philosophy and Its Others: Ways of Being and Mind (1990); Beyond Hegel and Dialectic (1992); Perplexity and Ultimacy (1995); Hegel’s God (2003); Art, Origins, Otherness: Between Art and Philosophy also in (2003); as well as Is There a Sabbath for Thought?: Between Religion and Philosophy (2005). He has also edited five books and published more than 100 articles and book chapters. His book The Intimate Strangeness of Being: Metaphysics after Dialectic appeared in 2012, the same year in which the William Desmond Reader also appeared. His new book, The Intimate Universal: The Hidden Porosity among Religion, Art, Philosophy and Politics, appeared with Columbia University Press 2016. He is Past President of the Hegel Society of America, the Metaphysical Society of America, and the American Catholic Philosophical Association. Nathan Eric Dickman (PhD, The University of Iowa) is Associate Professor of Philosophy and Religious Studies, and department chair, at Young Harris College. He researches in hermeneutic phenomenology, philosophy of language, and comparative questions in philosophy of religions. He has published numerous articles in Sophia,  Literature and Theology,  Religions,  The Humanistic Psychologist, and more. He teaches a breadth of courses, such as Existentialism, Muslim Journeys, Ethics, and the Historical Jesus. He recently served as the President of the Georgia

Contributors  377 Philosophical Society, and the philosophy of religion section Chair for the Southeast Commission for the Study of Religion. Janet Donohoe is Dean of the Honors College and Professor of Philosophy at the University of West Georgia. Her primary research interests are in phenomenology and hermeneutics. She is the author of several articles on phenomenology and place, as well as a book on Husserlian ethics and a book titled Remembering Places (Lexington Books, 2014). She recently edited a volume on Place and Phenomenology (Rowman & Littlefield, Int., 2017). Trish Glazebrook is Professor of Philosophy at Washington State University. She writes on Heidegger, ecofeminism, ancient philosophy, philosophy of technology, environmental philosophy, and climate change. Current research addresses climate impacts and adaptations by women subsistence farmers in Ghana. She also researches military use of drones, and other issues in military ethics. Michael Goldsby is an Assistant Professor of Philosophy at Washington State University. His primary research interests are in the areas of philosophy of science, epistemology, and practical ethics, but he sells himself as an applied philosopher with heavy pragmatic sympathies. More specifically, Goldsby works on applying philosophical insights to so-called “wicked problems.” One such wicked problem is maintaining food-energy-water security in the face of climate change, and he collaborates on a project (jointly funded by the USDA and NSF) aimed at solving that problem. Todd May is Class of 1941 Memorial Professor of Philosophy at Clemson University. He is the author of fourteen books, most recently A Fragile Life: Accepting our Vulnerability (University of Chicago Press). He has written extensively on recent French thought, particularly that of Michel Foucault, Gilles Deleuze, and Jacques Ranciere. Phillip E. Mitchell is Assistant Professor of English at the University of North Georgia. He earned his PhD from the University of Newcastle upon Tyne in 2015. He has published research in New Writing: The International Journal for the Practice and Theory of Creative Writing and creative work—nonfiction, fiction, and poetry—in several journals. Gregory S. Moss is currently an Assistant Professor of Philosophy at Chinese University of Hong Kong. He specializes in post-Kantian German philosophy, and has published in a variety of philosophical journals. Before completing his PhD on Hegel’s Logic of the Concept under Richard Winfield, he was a Fulbright Fellow with Markus Gabriel at the Rheinische Friedrich-Wilhelms-Universität Bonn. Robert Cummings Neville is Professor Emeritus of Philosophy, Religion, and Theology at Boston University, where he is also Dean Emeritus of

378 Contributors the School of Theology and of Marsh Chapel; in addition he has taught at Yale University, Fordham University, SUNY College at Purchase and SUNY University at Stony Brook. He has been president of the American Academy of Religion, the Metaphysical Society of America, the International Society for Chinese Philosophy, the Institute for American Religious and Philosophical Thought, and the Charles S. Peirce Society. He has published many books and articles, including most recently a philosophical theology trilogy: Ultimates (2013), Existence (2014), and  Religion (2015), as well as The Good Is One, Its Manifestations Many: Confucian Essays in Metaphysics, Morals, Rituals, Institutions, and Genders (2016), and Defining Religion (2018). Graham Priest is Distinguished Professor of Philosophy at the Graduate Center, City University of New York, and Boyce Gibson Professor Emeritus at the University of Melbourne. He is known for his work on nonclassical logic, metaphysics, the history of philosophy, and Buddhist philosophy. He has published nearly 300 articles—in nearly every major philosophy and logic journal—and seven books—mostly with Oxford University Press. Further details can be found at: grahampriest.net. Robert H. Scott is an Assistant Professor of Philosophy at the University of North Georgia. His research focuses on phenomenology and environmental ethics, and in recently published work he develops a phenomenological theory of ecological responsibility. Scott currently serves as the President of the Georgia Philosophical Society. J. Aaron Simmons is Associate Professor of Philosophy at Furman University in Greenville, South Carolina. He is the author of God and the Other: Ethics and Politics After the Theological Turn, co-author of The New Phenomenology: A Philosophical Introduction, and his edited books include Kierkegaard’s God and the Good Life and Kierkegaard and Levinas: Ethics, Politics, and Religion. Qingjie James Wang is Professor at the Department of Philosophy of the Chinese University of Hong Kong. He specializes in Heidegger’s Philosophy, Hermeneutics, East-West Comparative Philosophy and Chinese Philosophy.  His recent publications include Moral Affection and The Confucian Exemplary-Formative Ethics of Virtue (2017) and Heidegger and the Beginning of Philosophy (2015). Wang is also one of the major translators of Heidegger’s works in China. His translations include  A New Translation of Heidegger’s Introduction to Metaphysics (2015), Kant and the Problem of Metaphysics (2011), Introduction to Metaphysics (1996, co-translator) and Being and Time (1987, co-translator). George Wrisley is Associate Professor of Philosophy at the University of North Georgia. After spending many years working within the context of

Contributors  379 Wittgenstein’s later philosophy, in recent years he has focused on understanding the work of the 13th-century Japanese Zen Master Dōgen. Two central aspects of that work have been a) to bring Dōgen and Zen into conversation with Nietzsche’s work on suffering, and b) to critique understandings of enlightenment as some kind of language-transcendent mental state.

Index

abandonment 17, 34, 98 Abraham 35 – 36, 271 absolute nothingness 22, 26 – 27, 31, 301 – 302, 359, 366, 368 absurd 35 – 36, 60, 136, 198, 232, 301, 320 actuality-primacy 39, 104, 115 aesthetics iii, 7, 37, 213, 221, 259, 376; experience 41, 248, 251 – 257 affordances 134 – 141 Africa 34, 148 – 152, 156, 162 Agamben, Giorgio 39, 105 – 106, 120 – 123 agathon/Form of the Good 142, 343 agriculture 148 – 152, 162, 250 alienation 84, 91, 93, 98, 102, 107 Alienworld 168 – 176 ambiguity 121, 133 – 140, 143, 150, 181 Anaximander xi, 1, 11 Anselm, St. 38, 121 Apeiron xi, 11, 51 Aquinas, St. Thomas 109, 121 – 122, 343, 349, 357 Arendt, Hannah 33 – 34, 45 Aristotle 1 – 6, 9, 11, 14 – 17, 29 – 30, 39 – 40, 57, 105, 109, 115, 121, 161, 164, 182 – 186, 190 – 197, 216 – 217, 230, 264, 303, 342 – 343; metaphysics 9, 40, 57, 109, 121 – 123, 184 – 185; peripatetic 1, 3 astonishment 54 – 65; ontological 37 – 38, 51 – 53, 55, 62 – 64 ἀρχή 7, 42 Avalokiteshvara 317 – 318, 326 αὐτὸ καθ ̓ αὑτὸ (itself by itself) 1, 10, 13

Baroque music 222 – 223 becoming 2 – 3, 6, 14, 23 – 25, 27, 34, 53 – 56, 63, 65, 76, 111, 113 – 114, 118 – 120, 128, 345; self-becoming 52, 54 – 55, 59, 65 being 1 – 18, 23 – 34, 38, 41, 51 – 65, 71 – 72, 75, 85 – 87, 91 – 99, 105, 107, 111, 113, 118 – 122, 127 – 145, 175, 186, 194, 198, 201, 211, 217 – 218, 233, 235 – 238, 248, 269, 279 – 280, 289 – 307, 316 – 321, 333 – 335, 338 – 339, 343, 345, 349 – 354, 357 – 368, 376; determinate being 5 – 7, 15 – 16, 23, 25, 27, 52, 56 – 57, 62 – 63, 113; wild being 5 being-morally-moved 339 Bhagavad Gita 350 Bildung 225 black 176 – 179, 255, 359 Bley, Carla 224, 227 Bodhisattva 285, 312 – 316, 319 – 325; ideal 27, 41, 285, 309 – 311, 317 – 318, 326 bodies 17, 133, 177, 233, 250, 269, 331 – 336, 339 – 340, 347 body 10 – 14, 24 – 25, 27, 41, 70 – 71, 117, 134 – 139, 177, 198, 233, 241, 250, 288, 321, 330 – 341, 344 – 347, 356 body schema 134, 136 Buddhism 280 – 281, 285 – 286, 307 – 313, 317 – 321, 324 – 326; Buddha mind 350; Mahayana Buddhism 281, 309, 317, 328 – 329; Zazen 26, 315; Zen Buddhism 280, 309, 311, 314, 319 Cage, John 40, 222, 224, 227 Canaanites 349

Index  381 Carter, Joseph P. 29, 45 categories 4, 13, 16, 19, 37 – 38, 54, 62, 67 – 73, 79 – 80, 82, 106 – 107, 115 – 116, 119 – 120, 134, 201 – 204, 209, 242, 333, 363 cause 53, 72, 107, 134, 147, 149 – 150, 152, 160, 172, 174, 184 – 186, 270, 272, 284, 309, 311, 321, 325, 339, 342 – 343, 348, 350 – 351, 361 Cavarero, Adriana 34, 201, 204, 208 – 211 certainty 8, 31, 37, 152 – 157, 166, 212, 215, 219, 261 – 262, 265 – 266, 270 – 272; epistemic certainty 153, 265, 269 chaos (Χάος) xiv, 6, 9, 249, 254 Chaoskampf 9 climate adaptation 161 climate change 39, 148 – 156, 160 – 162, 199 – 200 climate denial, deniers 39, 149 – 150, 152 – 155, 157, 160 – 161, 165 climate models, modeling 149, 152 – 154, 157, 160 – 161, 165 community 44, 60, 121, 153, 178 – 179, 222, 236, 264, 271, 335, 338, 340 compassion/compassionate 309, 311, 316 – 328 completeness 8, 12, 20, 68, 70, 93, 172 comportment 39, 135 – 145 conatus essendi 56, 58, 63 Confucianism 42, 330, 333 – 334, 336 – 341, 351, 354; NeoConfucianism 334, 346, 351 – 352, 356 Confucius 333, 338, 345 conjunction 68, 114, 119 contingency 34, 38, 67, 69 – 70, 73 – 77, 105 – 108, 114 – 118, 122 – 123 contradiction 3 – 7, 14 – 15, 22, 24, 32, 35, 37, 75, 77, 84, 88, 90 – 93, 97, 102, 106 – 107, 111 – 122, 153, 210, 217, 248, 280 – 283, 286 – 289, 292 – 299, 302 – 304 conversation 9, 27, 34, 40, 226, 228, 245, 252, 284, 298, 309, 379 copula 119, 237, 239 cosmology 332, 342, 354, 357, 373; questions 15, 42, 357, 360 – 363, 366, 371 – 372 creatio ex nihilo 216

creativity 24, 315, 322 – 323, 327, 353 – 354 crisis 22, 27, 32, 84, 91 – 92, 99, 147, 149, 151, 156, 162, 164, 168, 178, 186, 190 – 194, 199 critique 4, 12, 19, 62, 67 – 68, 76, 81, 123, 145 – 146, 157, 170, 176 – 180, 193 – 194, 199 – 201, 206, 210, 243, 263 – 266, 328, 342, 379; renewal and critique 40, 168, 175 – 177, 179 – 180 curiosity 54 – 59, 62 – 64, 357 Daoism 42, 286, 301, 341, 356; cosmological 351, 353; Dao 15, 42, 287, 301, 332, 348, 350 – 355; Daodejing 346, 349 – 352, 355 – 356; Eternal Dao 350 – 352;Way (Tao) 347, 350 – 352, 354 Dasein 5, 15 – 16, 27 – 31, 34, 42, 138 – 141, 300 decision 77, 114, 142, 155, 193, 220 – 221, 253, 267, 271, 272 deduction 10, 19, 71, 73 – 75, 79 – 80, 82 – 83, 107, 115, 118, 192, 199, 334, 336, 339; genetic 68, 72, 78, 80, 82; metaphysical 68 – 69, 78; transcendental 68 – 70, 78 – 79, 83 dehiscence 233 – 234 Deleuze, Gilles 201, 377 delusion 287, 309 – 310, 314 – 321 dependent origination 25 Derrida, Jacques 23, 40, 44 – 45, 201, 204, 208 – 211, 241 – 246; deconstruction 22 – 23, 244, 246; différance 23, 204, 211; logocentrism 23; trace 23 Descartes, René 31 – 32, 109, 198, 337, 343, 354 determinacy 1, 4 – 22, 25 – 26, 33 – 38, 40, 51 – 53, 62 – 63, 67 – 68, 72, 74, 76, 79, 87 – 88, 91, 97, 115 – 116, 120, 129 – 132, 136, 173, 176, 180, 184, 188, 195, 215 – 220, 223 – 226, 261 – 262, 265 – 266, 269 – 272, 343, 366 determination 2, 4, 7, 9 – 12, 16 – 17, 23 – 24, 27, 33, 36 – 38, 43, 51 – 76, 79 – 80, 84 – 98, 107, 110, 119 – 120, 129 – 136, 139, 143, 188, 196 – 197, 215 – 218, 226, 232, 236 – 238, 261 – 265, 269, 272 – 274, 280, 294 – 295, 338 – 340, 343

382 Index diachronic 189, 199, 312 – 313, 325 dialectic 14, 17, 22 – 23, 35 – 36, 39, 52 – 53, 62, 75, 88, 104, 107 – 109, 114 – 115, 259, 373, 376 dialetheism 22, 289, 294, 375 differentia διαφορά 9 disjunction 60, 113 – 114, 120 Dōgen 22, 41, 280 – 281, 287 – 288, 306, 310 – 318, 321 – 328, 379 Dreyfus, Hubert 135 – 139, 142, 144 – 146 ecstasy 33, 84, 96 – 100, 102 – 103 Ekstase 96 emanation (απόρροια) 10 emptiness 25, 75, 241, 279, 282, 287, 302, 306, 313, 315, 318, 323, 327, 350 Enlightenment 25, 284 – 285, 288, 309 – 311, 314 – 324, 379 epistemology xiii, 6, 17, 24, 51, 57, 79, 85 – 88, 91, 99, 127, 132 – 133, 153 – 155, 160, 185, 189, 203, 208, 261 – 262, 265, 267 – 272, 279, 315, 320, 353, 377; epistemic imperatives 228, 231 equivocal 52, 60 – 61, 136 – 137 equivocity 52, 60 – 64 Erkenntnis 4, 238, 373 ethics iii, xii, 24, 34, 37, 77, 109, 179, 181 – 183, 190 – 191, 195 – 200, 208, 243 – 244, 266, 268, 327, 338, 340; Confucian 24, 41, 330, 339 – 341 evasiveness 8, 27 ex nihilo nihil fit 15, 342, 368 existentialism 14, 21 – 22, 25, 28 – 34, 37 – 38, 41, 77, 136, 138, 192, 194, 233, 238, 240, 248, 261, 269 – 271, 298, 335, 353; determination 262, 265, 269, 272 – 274 experience 12 – 13, 19 – 21, 33, 36, 38 – 41, 55, 67 – 74, 79, 81, 93 – 97, 100, 128 – 134, 138, 144, 161, 168 – 178, 189, 191, 197 – 198, 201 – 204, 215, 223, 229 – 230, 233 – 238, 242, 247 – 248, 251 – 257, 263, 268, 281 – 282, 286, 312 – 314, 320 – 322, 325, 332, 334, 338 expression 137, 208, 211, 233 extending oneself to others 336, 338 ἐντελέχεια (Being-at-work staying itself) 3, 15, 29

fecundity 351 Fichte, Johann Gottlieb 38, 67 – 83; Wissenschaftslehre 69 – 74, 79 – 83 food security 34, 148 – 152, 156, 162, 166 form (εἶδος) 1 – 3, 10, 16, 23, 37, 71, 80, 92, 128 – 129, 132, 134, 136, 286, 302, 343 Frageverhalt 237 – 238 freedom 17 – 18, 28, 31 – 36, 53, 67 – 78, 81, 90 – 100, 218 – 219, 225, 244, 307; eternal freedom 90, 93; free thinking 15, 36, 38, 84, 92 – 100 function argument 182, 196 Gabriel, Markus iii, 21 – 22, 44, 46, 89, 91, 94, 102 – 103 Gadamer, Hans-Georg 37, 40, 215, 219 – 222, 225 – 227, 230, 235 – 236, 238 – 246, 289 Galileo 164, 193 – 194, 268 genius 37, 315, 324 givenness 20, 53, 64, 117, 130, 144, 171 – 172, 175 – 176, 187 – 188, 191 God 10, 12, 35 – 38, 44, 63 – 66, 98, 100, 102, 118, 121, 198, 216 – 219, 236, 240, 244, 262, 269 – 272, 291, 338, 340, 343, 345, 349, 357 – 360, 366, 368, 371 – 372, 376, 378 goddess 284, 289 Gödel’s incompleteness theorem 8 Goethe 316 – 319, 324 Golden Rule 336, 341 graded arrangement 338, 340 Great Ultimate (T’ai-chi) 352 – 353 Grundsatz der Bestimmbarkeit (Principle of Determination) 4, 11 Grundsatze der durchgängingen Bestimmung (The principle of thoroughgoing determination) 11 – 12, 43 Hartshorne 343 Heaven (T’ien) 338, 345 – 347, 350, 355 – 356 Hee-Jin Kim 314 – 318, 326 Hegel, G.W.F. 19, 21, 27, 33 – 35, 38, 51, 97, 99, 102, 104 – 122, 242, 279; anxiety 31, 59, 176, 244, 300; Aufhebung (dialectic) 7, 14, 16 – 17, 22 – 23, 35, 39, 52 – 53, 62, 74 – 75, 88, 104 – 109, 259, 373, 376; authenticity/inauthenticity 28, 30, 146; conscience 31, 244; Hegelian negativity 53 – 54, 64 – 67, 73, 77 – 81

Index  383 Heidegger 9, 12, 20, 22, 25, 27, 29 – 31, 33 – 34, 38 – 39, 41 – 42, 54, 105 – 106, 109, 120 – 121, 127, 138 – 146, 149, 161, 164 – 165, 209, 231 – 233, 243, 280 – 281, 289 – 306, 357, 359, 366 – 368, 371 – 372, 375; Holzweg 28; Phenomenologie des Geistes 17; Science of Logic (Wissenschaft der Logik) 5, 15 – 17, 39, 46, 68, 79, 83, 103, 300; Sein Zum Tode (Being-toward-death) 28 – 29, 31, 45 Heisenberg’s Uncertainly Principle 8 Henry, Michel 37, 41, 262, 266 – 274 Heraclitus 2, 216 hermeneutics xii, 40, 261, 230, 244 – 245 higher types 315 – 316, 321, 324 higher-order we 178 – 179 Hirsch, E.D. 223, 227 home 21, 55, 62, 93, 168 homeworld 21, 168, 170 – 178 hooks, bell 322 hope 89, 254, 262, 272 – 274, 328 horizon 20 – 21, 28, 30, 39 – 40, 44, 46, 127 – 144, 168 – 180, 182 – 190, 193 – 199, 208, 229 – 232, 239; external 130 – 132, 143, 170 – 171, 185, 190, 195, 198 – 199; of indeterminacy 39 – 40, 182 – 190, 193, 195 – 198, 239; internal 130 – 132, 170 – 171, 190; world 21, 34, 39, 127 – 132, 134 – 136, 138 – 144 168 – 179, 182 – 190, 193 – 199, 353 human function 182, 196 – 197 human persons 217 – 218 Husserl, Edmund xi-xii, 13 – 14, 17, 19 – 21, 27, 31 – 32, 39 – 40, 127 – 139, 143 – 145, 168 – 172, 176 – 180, 186 – 199, 229, 236, 239, 241 – 242; Crisis of the European Sciences 27; descriptive phenomenology 14; ὁρίζων 20 – 21; epoche 17 – 18, 129, 138, 191 – 192, 198, 233, 235, 237; genetic phenomenology 21, 40, 44, 168, 170, 174 – 175, 177 – 178, 181, 187 – 189, 198 – 200; natural attitude 17 – 21, 128 – 129, 138, 187, 198, 229, 232, 235, 237; noesis and noema 19; Philosophy as Rigorous Science 18, 144; transcendental ego 19 Hutcheon, Linda 255 – 259

I-hood 71, 74 idea 12, 21, 40, 52, 74 – 76, 80 – 81, 89, 107, 137, 143, 168 – 169, 172, 186 – 187, 198, 201, 211, 224, 233, 263 – 264, 285, 295 – 296, 298 – 299, 306, 309 – 311, 325, 332 – 340, 343, 350, 354, 370 – 371 identity 26, 33, 70, 76, 80, 85, 98, 106 – 107, 111 – 113, 118 – 120, 129, 132 – 134, 188, 253, 262, 270, 274, 281, 286, 311, 318, 323, 327, 362 – 365, 370 inclusion 108, 120 incompleteness 8, 23, 29 – 30, 74, 80, 113 incomprehensible 34, 74 – 77, 87 – 89, 100, 136 inconceivability 7, 38 indefinability 7 indeterminacy iii, xi-xv, 1 – 44, 51 – 54, 59, 64 – 68, 75 – 77, 84, 90, 92, 97, 100, 111, 113 – 116, 120, 127 – 139, 142 – 144, 148 – 149, 152 – 153, 156 – 157, 160 – 162, 168 – 190, 193, 195 – 199, 201, 215 – 219, 222 – 226, 228 – 239, 243, 248, 253, 261, 265, 269 – 274, 279, 304, 310 – 315, 320, 322, 325, 338, 342 – 344, 348 – 350, 353 – 354, 362; absolute 5, 8 – 12, 17, 20 – 22, 27, 84, 88 – 91, 93, 96 – 97; places of indeterminacy (unbestimmtheitsstellen) 224; relative 5, 8 – 9, 20, 25, 84, 88, 97, 129; void indeterminacy 52 indeterminate xi-xiv, 1 – 42, 51 – 54, 59, 62 – 64, 68, 75, 77, 79, 88 – 97, 100 – 101, 111 – 113, 120, 122, 127 – 135, 138, 142, 168, 171 – 176, 179 – 180, 187, 189, 198, 201, 216 – 226, 229, 231, 235 – 239, 248, 252 – 257, 279 – 280, 295, 300, 310, 314, 330, 336 – 340, 343, 345, 348 – 353, 358 – 366, 371 indifference 18, 77, 88, 98, 107, 117, 162, 256 ineffability 7, 40, 201, 204, 280, 282, 284, 289, 293, 299 – 302, 307 infinite 7 – 8, 80, 83, 87, 90, 97; bad infinite 80, 83, 97; perfect/ perfection 3, 7 – 8, 156 – 159, 217, 282; quantitative infinite 7 – 8; true infinite, 87, 90, 97 Ingarden, Roman 40, 146, 223 – 224, 227

384 Index intellectual intuition 38, 43, 76, 80 – 81, 96, 98 intentional implications 128, 130 – 131, 134, 138, 140, 142 – 143 intentionality 19 – 21, 34, 56, 134 – 139, 142, 145, 168, 170, 189, 233 – 234, 237, 241, 272; of experience 189; motor intentionality 25, 134 – 139 intersubjectivity 144, 189, 199, 230 intimateness 24, 330, 333 – 339, 341 intuitive reason 183, 186, 190 – 195 invariable objects 183 – 185, 195 – 196 IPCC 148, 152, 153, 154, 156, 160, 162, 163, 164 Irigaray, Luce, 234, 244 Jacobi 33, 103 jazz 37, 215, 222 – 227 Jen 346 judgment 17 – 18, 26, 34, 36 – 37, 68 – 69, 71, 77, 90 – 91, 95, 130, 147, 187 – 189, 199, 207 – 208, 211, 220 – 221, 225, 229, 230 – 239, 244, 342, 359 justice 39, 148, 150 – 152, 162, 165 – 166, 211, 261, 321, 323, 326 – 327 Kant, Immanuel iii, xi, 3 – 5, 8, 10 – 13, 16, 19, 38, 42 – 43, 46, 57, 62, 65, 67 – 70, 73 – 74, 77 – 83, 91, 94, 107, 121 – 123, 135, 142, 145 – 147, 220 – 221, 241 – 242, 244, 342, 357, 373, 375, 377 – 378; Critique of Pure Reason: 4, 46, 67 – 68, 78, 83, 121, 123, 244, 342, 373 Keats, John 248 Kierkegaard, Søren 46, 262, 275; despair 35; leap of faith 35 – 36 knowledge 3, 4, 10, 13, 14, 16, 18, 20, 22, 24, 32, 36, 38, 41, 46, 74, 76, 78, 82 – 83, 85 – 88, 90 – 94, 96, 98 – 100, 102, 129, 131, 143, 160 – 161, 165, 177, 182 – 183, 185, 186, 190 – 196, 219 – 220, 225, 234, 240, 255 – 256, 261 – 262, 266, 268 – 269, 272, 330, 334, 336, 341, 354, 356 Kyoto School xi, 22, 25, 44, 302 Laozi 307, 345, 355 – 356 Leibniz, Gottfried Wilhelm 39, 107, 343, 357, 366

Levinas, Emmanuel 22, 40, 136, 145, 147, 199 – 202, 204, 208 – 212, 229, 234, 240 – 241, 243 – 245, 273, 275, 378 Li 346, 348, 355 Lifeworld 170 – 173, 175 – 176, 191, 193 – 195, 229, 242 limitation 4 – 5, 24, 32, 42, 68, 72, 74, 90, 117, 120, 127, 202, 340, 343 logic 3 – 7, 15 – 17, 19, 22, 27, 38 – 39, 52 – 54, 67 – 71, 73, 75, 78 – 83, 91, 102, 104 – 110, 113, 118, 120, 122 – 123, 127, 130, 136, 142, 149, 154 – 155, 169, 187, 189, 200, 215, 219 – 220, 228, 231, 234, 243 – 244, 262, 265, 270, 273, 279, 292 – 294, 297, 300, 302 – 304, 306 – 308, 325, 339, 358 – 359, 363, 365, 373; modal logic 39, 67 – 68, 104 – 110, 113 – 115, 118 – 123; paraconsistent logic 22 love 25, 36, 46, 179 – 180, 244, 247, 248, 250, 255 – 256, 264, 272, 312, 314, 326, 330, 333, 335 – 340, 349; love without gradation 兼 愛 337 Lyotard, Jean-François 22 Marion, Jean Luc 38, 46 McKenna, Tony 247, 254, 259 McPartland, Marian 225 meaning, xiii, 9, 19 – 20, 25, 33 – 34, 37, 39, 42, 53, 97, 130 – 132, 134 – 139, 141 – 143, 145, 151, 154, 169, 173, 176, 178, 182 – 190, 192 – 197, 199, 222 – 223, 228 – 236, 238, 240, 242, 244, 245, 248 – 252, 254 – 255, 264, 266 – 269, 284 – 285, 290, 292, 294, 298, 305, 310, 319, 325, 331, 333 – 335, 341, 343, 356, 376 Megarians 105 Merleau-Ponty, Maurice 8, 20, 34, 39, 127, 133 – 139, 141, 143 – 147, 198, 233, 240 – 245 meta-metaphysics 85 – 87, 89, 99 – 100, 102 metaphysics xiii, xiv, 6, 9, 15, 45 – 46, 51, 54, 57 – 58, 62, 65, 67 – 70, 78 – 79, 85 – 89, 91, 93 – 94, 99 – 102, 105, 117, 121 – 123, 129 – 130, 132 – 133, 136, 139, 144 – 147, 184 – 185, 200, 209, 267, 279, 289, 290 – 299, 306 – 308, 314, 318, 343, 352, 355 – 357, 360, 362, 366, 368, 371, 373

Index  385 Meyer, Michel 231, 245 model selection theory 157, 159, 164 motion (Kinesis) 23, 29 – 31, 65, 193, 216, 254 motor intentionality 134 – 136, 138 – 139 multiplicity 123, 170, 190, 196, 198, 328 mysticism 92, 100, 102 – 103 Myth of the Given 14 Nagarjuna 25; Fundamental Wisdom of the Middle Way 281 – 283, 304 Nancy, Jean-Luc 104, 123 negation 1, 4 – 7, 10, 14 – 15, 17 – 18, 25 – 27, 33, 42, 53 – 56, 59, 62, 65, 75, 80, 87 – 88, 93, 104, 112 – 113, 119, 239, 301 – 302, 370 negative capability 248, 254 negativity 6, 51, 53 – 57, 61, 64 – 65, 104 – 105, 107, 117 – 119, 238, 302 New Critics 248 Nietzsche, Friedrich 46, 83, 216, 227, 259, 261, 328, 341; The Birth of Tragedy 251, 259; Übermensch 24 Nirguna Brahman 350 Nishida, Kitaro 42, 357, 359, 366 – 368, 371 – 373 Nishitani, Keiji 22, 25 – 27, 33, 44, 46, 281, 301 – 303 non-being 2, 5 – 6, 25 – 26, 30, 287, 297, 302, 349 – 354 noncontradiction 2, 4, 7, 22, 88, 280, 293 – 297, 306 nondual 285, 313 – 315 non-positional 6 – 10, 14 – 18, 22 – 23, 25 – 26, 31 – 32, 36 – 37 non-predicative 187, 189 non-white 177 normativity 34, 40, 134 – 135, 138, 142, 146, 202 – 206, 208, 210 – 211, 376 nothingness 6, 22, 26, 27, 31, 46 – 47, 147, 281, 299 – 303, 306, 308, 324, 359, 366, 368 Noumenon 13, 16, 43, 241 νοῦς 2, 15 omnis determinatio est negatio 5 Ontology 24, 46, 103, 119 – 120, 137 – 138, 145, 183, 185, 202, 208 – 209, 240, 244, 325, 330, 371, 375; ontological horror 61, 65 optimality 135 overdeterminacy 37, 51 – 54, 62

pain 61 – 62, 71, 73, 230, 247, 256 – 257, 310, 313 – 315, 319 – 322, 324 – 328, 368 pan-moralism 339 paradox 6, 9, 13, 26, 34 – 36, 41, 44, 100, 106, 112 – 113, 115, 121, 256, 259, 291 – 292, 294 – 295, 303, 305, 308, 313 Parmenides 2, 10, 60 perception 19, 128, 130 – 133, 135 – 136, 144, 146 – 147, 168, 170, 172, 190, 198 – 199, 245, 268, 325, 353 – 354 perceptual phenomenology 130, 132 – 133, 138, 169, 171, 197 perplexity 59 – 65, 264, 376 philosophy: global philosophy 354; philosophic wisdom 14, 40, 182 – 184, 186, 190 – 191, 195 – 196, 264, 273 Phronesis 183, 197, 217 Phronimos 217 Picasso, Pablo 221 Plato xiii, 1 – 3, 10 – 11, 37, 42 – 43, 46 – 47, 57, 61, 65 – 66, 142, 208 – 209, 233, 241, 245, 259, 264, 291, 305, 307 – 308, 316, 342 – 343, 345, 349, 356; allegory of the cave 61 – 62; form of the good 343; Parmenides 10, 46, 356; Republic 3, 245, 342 – 343; Seventh Letter 10, 47 Plotinus 10, 343, 351; Enneads 343; Neo-Platonism 305, 343 porosity 38, 41, 55 – 59, 61, 63 – 65, 376 positionality 8, 14, 16, 23, 238 possibility vii, 3, 5, 7, 9, 11 – 13, 15, 20, 22, 28 – 32, 38 – 39, 53, 56, 63, 68, 70 – 71, 73, 79 – 80, 85, 90, 104 – 123, 131, 136, 138, 140, 149, 161, 168, 172, 175, 177, 179, 185, 198, 202, 205, 215, 262, 298 – 99, 361, 371, 375; absolute 105, 108, 115, 118; actualized 105, 108, 114 – 115, 118; formal 108, 111 – 118, 120, 122; itself 114 – 120; negative side of (not to be) 104 – 106, 110, 112 – 113, 115 – 120; positive side of (to be) 104 – 106, 110, 112 – 113, 115 – 116, 118, 120, 122; qua possibility 113, 115 – 118, 122; real, 108, 113 – 122; unactualized 39, 105, 108, 114 – 115, 118 Possible Worlds 139, 142, 173, 175 poststructuralism 22

386 Index potentiality 29 – 31, 33, 105, 113, 121, 123, 349 practical wisdom 182 – 184, 186, 190 – 192, 195 – 196, 313 practice 3, 22, 26, 36, 40 – 41, 85, 89, 100, 131, 136, 143, 147, 155, 170, 192 – 196, 203 – 206, 208, 210, 212, 216 – 217, 220, 222, 224 – 225, 227, 243, 268, 271, 287, 309 – 310, 315, 319, 320 – 324, 327, 332, 334, 336, 338, 341, 346, 377 predication 4, 12, 87, 92 Pre-givenness 172, 176 Pre-predicative 242 Presupposition 17 – 19, 70 – 71, 73 – 77, 81, 107, 176, 195, 338 Priest, Graham 22 Principle of Determinability 4, 71 – 72, 75, 79 Probability 149, 152 – 154, 156, 247 Proclus 11, 308 Project 34, 134 – 140, 143, 145 – 146, 148 Pseudo-Dionysius 10 Pyrrhonian skeptics 18 παλίντροπός (back turning) 2 Qi 344, 346, 348 question quid juris 68, 78, 80 question quid facti 68, 70, 73 – 74, 78, 80 Questioning 22, 40, 51, 61, 155, 170, 207, 228 – 245, 263 – 264, 267, 373 race 168, 173, 177 racism 21, 40, 177 – 179 reason xiii, 4, 12, 20, 26 – 28, 34 – 37, 46, 65, 67 – 68, 70, 74, 76 – 78, 80 – 81, 83, 85 – 87, 93, 99, 103, 107, 121 – 123, 132 – 133, 136 – 137, 147, 154 – 155, 157, 165, 182 – 188, 190 – 193, 195 – 198, 203, 205, 219 – 221, 231, 244, 293, 302, 307, 325, 334, 338, 341 – 343, 346, 349 – 350, 358 – 363, 373 reductionism 201 – 202 Ren 335, 336, renewal 40, 168, 175 – 180, 187 responsibility iii, 77, 142, 148, 150, 152, 160 – 161, 177 – 180, 199 – 200, 211, 222, 232, 234, 239 – 240, 243, 318 – 320, 378 Ressentiment 64, 324

rhapsody 68 – 69, 71, 73 – 76 Ricoeur, Paul 229, 236, 245 risk 34, 58, 85, 162, 209, 234, 261 – 262, 264 – 266, 269, 271, 273 – 74 romanticism 37, 46 rule following 203, 206, 220, 225, 339 Sartre, Jean-Paul 31, 47, 147, 245 Schelling, F.W.J. xi, 18, 33 – 34, 36, 38, 44, 67 – 68, 76 – 77, 81 – 91, 93 – 103, 308, 375 Schlegel, Friedrich 37 scientific knowledge 22, 74, 92, 94, 98, 100, 143, 156, 160 – 61, 183, 186, 190 – 196, 219, 266, 268, 334 sedimentation 168, 173 – 174 self 90, 95 – 98, 142, 211, 234, 254, 287, 310 – 312, 315, 318, 321, 330 – 341, 347 self-determination xiii, 52 – 57, 59, 62, 64 – 65, 79, 98 Senauke, Alan 318, 328 Shangdi 345 Shen ti 身體 331, 333 – 34 sincerity (cheng) 347 skepticism 18, 62, 65, 68, 79, 82, 87, 89, 94, 152, 375 Spinoza, Baruch 107, 109 spontaneity 15, 37, 215 – 216, 222, 224 – 225, 342, 353 STEM 260, 262, 265, 267 – 269, 273 – 274 Stiegler, Bernard 105, 123 Stoner 37, 41, 247 – 259 subject (sentential) 229, 234 – 237 sublime xiv, 37, 66, 246 substance 80, 106, 108, 115 – 116, 118, 121, 183 – 85, 195, 248, 250, 256, 311, 318, 331, 334, 344, 354 suffering 65, 148, 150, 162, 310, 313 – 327, 379 summons 72, 81 – 82 synchronic 312 – 313, 325 system 8, 16, 24, 27, 34, 38, 52, 67 – 68, 70 – 71, 73, 75 – 77, 81 – 83, 85, 87 – 89, 91, 94, 98, 100, 102 – 103, 108, 123, 134, 148, 150, 152 – 153, 155 – 157, 159 – 161, 177 – 179, 190, 199, 202, 232, 266, 268, 274, 295, 310, 316, 318, 323, 331 – 332, 338 – 339, 344, 346, 352, 354, 375

Index  387 tact 219, 220, 225 taijiquan 344, 354 taste 37, 201, 219, 221 – 222, 225 – 226, 272 teleology 29, 64, 269 thaumazein 57, 64 theology 10 – 11, 198, 219, 246, 356, 368 thing-in-itself 12 – 13, 20, 241 Tillich, Paul 231, 245 – 246, 343 Tohu va vohu 216 totality 7, 11, 22, 28 – 30, 62, 64, 72, 76, 106, 109, 111 – 112, 114 – 115, 118 – 120, 122, 138, 140, 144 – 145, 147, 245, 286, 331, 358, 360 – 361, 370 tradition 1, 4, 5, 9 – 11, 13 – 14, 22, 25, 29 – 30, 40 – 41, 44, 53, 67, 69 – 71, 73, 77, 86, 100, 104 – 106, 108, 110, 114 – 115, 118, 134, 161, 165, 168, 172 – 173, 178 – 179, 186, 208 – 209, 217, 219, 225, 232, 244, 280 – 283, 285, 286, 289, 304, 310, 317, 331 – 333, 336 – 338, 340, 344 – 345, 348 – 354, 358, 360, 371 transcendental ideal 4, 12 trust 262, 271 – 272, 274 truth verification 104 – 106, 110, 122 Übergottheit 95, 100 – 101 ultimate reality 41, 74, 76, 280 – 282, 284, 286, 289, 299, 301 – 304, 306 uncertainty 7 – 8, 39, 59, 68, 149 – 150, 152, 154 – 157, 162, 215, 229, 261, 270, 272 undecided 68, 77 undifferentiated 95 – 96, 98, 279, 343 unity 16, 19, 37, 69, 75, 80, 86, 89, 96, 99, 111 – 113, 116 – 117, 122, 133 – 134, 140, 179, 185 – 186, 236, 238, 248, 314, 332, 354 – 356 univocal 51 – 52, 54, 65, 223 unknowability 7 unmoved mover 184 – 186, 195, 197, 216

vagueness xiv, 113, 307 – 308 validities 169 – 170 value 17, 34, 38, 41, 67 – 68, 76 – 77, 134, 158, 160 – 161, 163, 165, 168, 180, 183 – 184, 186, 194, 229, 253, 260, 264 – 265, 268 – 269, 273, 309 – 311, 316 – 317, 319, 322 – 325, 330, 336 – 338, 340, 347 – 348, 355 variable objects 184, 185, 191 Vaughan Williams, Ralph 223, 227 Via Negativa 10 voice 34, 201 – 206, 208 – 221, 242, 244 Wang Bi 301, 307, 350 – 351 Wang Yangming 333 Weil, Simone 272, 275 Whitehead, Alfred North 345, 354, 356 whiteness 177 – 178, 180 Will xiii, 26, 68, 77 – 78, 91, 98, 179 Williams, John 37, 247, 258 – 259 Wing-tsit Chan 351, 356 Wissen 84, 90 – 95, 99 – 100 Wissenschaft 14 – 15, 17 – 18, 46, 69 – 71, 73 – 74, 79 – 85, 92, 103, 123, 227 Wittgenstein, Ludwig 212, 246 wonder 36 – 37, 51, 54 – 65, 136, 143, 177 – 178, 229, 249, 253, 326, 357, 363, 368, 373 Wright, Dale S. 309, 328 – 329 Yahweh 349 Yang 257, 330, 332, 344 – 346, 348, 352 – 354 Yijing 345, 351, 354 Yin 332, 344 – 346, 348, 352 – 354 Zeus 345, 349 Zhongyong (The Doctrine of the Mean) 347 Zhou Dunyi 350, 352, 354 Zhuangzi 6, 42, 47 Žižek, Slavoj 104, 123, 231, 246