You are on page 1of 556

The Oxford Handbook of Time in Music

Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468768 by National Science & Technology Library user on 26 May 2023
FRONT MATTER

Copyright Page 
https://doi.org/10.1093/oxfordhb/9780190947279.002.0003 Page iv
Published: December 2021

Subject: Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

p. iv

Oxford University Press is a department of the University of Oxford. It furthers

the University’s objective of excellence in research, scholarship, and education

by publishing worldwide. Oxford is a registered trade mark of Oxford University

Press in the UK and certain other countries.

Published in the United States of America by Oxford University Press

198 Madison Avenue, New York, NY 10016, United States of America.

© Oxford University Press 2022

All rights reserved. No part of this publication may be reproduced, stored in

a retrieval system, or transmitted, in any form or by any means, without the

prior permission in writing of Oxford University Press, or as expressly permitted

by law, by license, or under terms agreed with the appropriate reproduction

rights organization. Inquiries concerning reproduction outside the scope of the

above should be sent to the Rights Department, Oxford University Press, at the

address above.

You must not circulate this work in any other form

and you must impose this same condition on any acquirer.


Library of Congress Cataloging-in-Publication Data

Names: Do man, Mark, editor. | Payne, Emily, 1985– editor. | Young, Toby, 1990- editor.

Title: The Oxford handbook of time in music / edited by Mark Do man, Emily Payne, and Toby Young.

Description: New York, NY : Oxford University Press, 2022. |

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468768 by National Science & Technology Library user on 26 May 2023
Includes bibliographical references and index. |

Identi ers: LCCN 2021029384 (print) | LCCN 2021029385 (ebook) |

ISBN 9780190947279 (hb) | ISBN 9780190947293 (epub)

Subjects: LCSH: Time in music. | Musical meter and rhythm. | Music—Psychological aspects.

Classi cation: LCC ML3850 .O94 2022 (print) | LCC ML3850 (ebook) | DDC 781.2/2—dc23

LC record available at https://lccn.loc.gov/2021029384

LC ebook record available at https://lccn.loc.gov/2021029385

135798642

DOI: 10.1093/oxfordhb/9780190947279.001.0001

Printed by Sheridan Books, Inc., United States of America


The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468803 by National Science & Technology Library user on 26 May 2023
FRONT MATTER

Acknowledgements  CP.P2

Published: December 2021

Subject: Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

MOST of the chapters in this volume developed out of papers presented at Making Time in Music, a conference CP.P3
at the Faculty of Music, University of Oxford in September 2016. The conference itself was made possible
through an Early Career Fellowship awarded to Mark Do man at the University of Oxford (2014–17) by the
Leverhulme Trust. The book, therefore, could not have been written without the Trust’s support.

The editors want to thank all the contributors for their illuminating scholarship on time in music: the CP.P4
volume presents as broad and rich a set of writings as we could have wished for. Our thanks also to our
colleagues at Oxford University Press—to Suzanne Ryan, for her enthusiasm for the volume in the initial
stages of the writing, and latterly to Lauralee Yeary and Norm Hirschy, for their help and support as we saw
the volume through to publication.

p. x Mark Do man, Emily Payne, and Toby Young CP.P5


The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468810 by National Science & Technology Library user on 26 May 2023
FRONT MATTER

List of contributors 
Published: December 2021

Subject: Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

Chloё Alaghband-Zadeh, Lecturer in Ethnomusicology, University of Manchester.

Nathan C. Bakkum, Associate Professor of Music, Columbia College Chicago.

John C. Bispham, A liated Researcher, Centre for Music and Science, Faculty of Music, University of
Cambridge.

Alexander E. Bonus, Visiting Assistant Professor of Music, Vassar College.

Andrew Bowie, Professor Emeritus of Philosophy and German, Royal Holloway, University of London.

Andrew Bowsher, Independent Scholar of Anthropology and Music.

Ryan D. W. Bruce, Instructor, University of Guelph.

Anne Danielsen, Professor of Musicology, RITMO Centre for Interdisciplinary Studies in Rhythm, Time
and Motion, University of Oslo.

Mark Do man, Lecturer in Music Psychology, University of She eld.

Rolf Inge Godøy, Professor of Musicology, University of Oslo.

Mark Gotham, Professor für Musiktheorie, Technische Universität Dortmund.

Anthony Gritten, Head of Undergraduate Programmes, Royal Academy of Music, London.

Kristina Knowles, Assistant Professor of Music Theory, Arizona State University.

Lawrence Kramer, Distinguished Professor of English and Music, Fordham University.

Juan M. Loaiza, Independent Scholar.

Nathan Mercieca, Independent Scholar.

Landon Morrison, College Fellow in Music, Harvard University.


Emily Payne, Lecturer in Music, University of Leeds.

Rainer Polak, Research Fellow, Max Planck Institute for Empirical Aesthetics, Frankfurt am Main.

Jonathan Roberts, Independent Scholar.

p. xii Floris Schuiling, Assistant Professor in Musicology, Utrecht University.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468810 by National Science & Technology Library user on 26 May 2023
Jonathan Still, Independent Scholar.

Chris Stover, Senior Lecturer in Music Studies and Research, Queensland Conservatorium, Gri th
University; Research Fellow, RITMO Centre for Interdisciplinary Study in Rhythm, Time and Motion,
University of Oslo.

Renee Timmers, Professor of Music Psychology, University of She eld.

Samuel Wilson, Tutor in Music Philosophy and Aesthetics, Guildhall School of Music and Drama, London;
Lecturer in Contextual Studies, London Contemporary Dance School; Independent Researcher.

Maria A. G. Witek, Senior Birmingham Fellow, Department of Music, University of Birmingham.

Toby Young, Leverhulme Early Career Fellow, Guildhall School of Music and Drama, London.
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468815 by National Science & Technology Library user on 26 May 2023
FRONT MATTER

About the Companion Website  CP.P6

https://doi.org/10.1093/oxfordhb/9780190947279.002.0007 Pages xiii–xiv


Published: December 2021

Subject: Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

www.oup.com/us/ohtm CP.P7

Oxford has created a website to accompany The Oxford Handbook of Time in Music. Material that cannot be CP.P8
made available in a book, namely audio and video clips, is provided here. The reader is encouraged to consult
this resource in conjunction the chapters. Examples available online are indicated in the text with Oxford’s
p. xiv symbol .
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
CHAPTER

1 Introduction  C1

Mark Do man, Emily Payne, Toby Young

https://doi.org/10.1093/oxfordhb/9780190947279.013.30 Pages 1–22


Published: 08 December 2021

Abstract
The work of this introductory chapter is twofold; rst, to provide a brief historical overview of the
changing nature and conception of musical time over the last 2,000 years, and second, to set out the
arc of the volume through detailing the central points of each chapter. While the individual pieces of
writing bring vital and varied perspectives from musicology, ethnomusicology, philosophy,
psychology, and sociocultural work, what unites them is their attention to music of the modern period,
with a strong focus on the multiplicities of contemporary practice, while also pointing to their
nineteenth-century antecedents. In introducing the main themes of the book, the introduction calls
attention to the burgeoning scholarship on time in music, ranging between the immediate feelings and
socialities of being in time with others and the broader imaginings of the cultural politics of time in
music.

Keywords: time, timing, temporality, rhythm, contemporary music, performance, composition,


interdisciplinarity
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

THE relationship between time and music is endlessly complex. As a ‘temporal art’, music’s interactions C1.P1
with—and representations of—the ways in which we experience time have o ered composers, performers,
and listeners throughout history profound insights into the human capacity, perhaps need, to delineate and
quantify our daily lives. As philosophers tell us, time itself may not be perceivable, but music allows us to
infer things about time in a way that many other aspects of human life cannot. Through its symbiotic
relationship with temporality, music is a orded a unique position in society as a place of external re ection
on or experimentation with the human experience of time.

Making music of whatever style or genre invites the actor to confront and mediate their experiences of the C1.P2
past and their expectations of the future on multiple scales by o ering them a ‘grid of musical space’ to
re ect on time and give it meaning and signi cance. As Suzanne Langer notes, ‘music makes time audible,
and its form and continuity sensible. […] It creates an image of time measured by the motion of forms that
seem to give it substance, yet a substance that consists entirely of sound, so it is transitoriness itself’ (1953:
110). In other words, as music unfolds in time it necessarily acts as a commentary on the temporal space in
which it exists, but it also generates its own time, creating particular temporalities outside of the physically
measurable time with the potential to shape or even suspend time for the performer or audience.

Of course, trying to untangle this complex relationship between music-making and the temporal C1.P3
imagination of geographically and historically diverse communities inevitably presents problems.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
Experiencing a Western classical performance of chamber music is clearly imbued with a di erent cultural
system of temporal representations and values from those of an Indonesian gamelan ensemble, and our
models of theorizing these experiences must re ect this complex relationship. This is achieved through no
single form of investigation, and we hope that the strength of this volume, beyond the individual
contributions, is to bring such varied perspectives into a rich dialogue.

p. 2
Musical Time: A Brief History C1.S1

While the main focus of the book is on contemporary music-making, the myriad ways in which music and C1.P4
time have interpenetrated throughout history provides a valuable frame of reference for the discussions in
this volume. Even in the earliest examples of music-making, cultural thinking about time was crucial to
understanding and contextualizing musical practice. For many ancient civilizations—including Assyria and
Babylonia, Classical India, the Hebrew tribes, China, Greece, and the Roman Empire—time was considered
to be a mythic force, an ‘all-embracing suprapersonal plane of temporality without which there could have
been no meaningful grounding in existence’ (Gross 1985: 4). For these archaic societies, the idea of a
temporal expanse was not considered as a passive void, but rather a static, cosmic object, from which all
earthly experience of normal time was born. This eternity was considered sacred and transcendental.
Philosophers like Boethius (and later Augustine) describe it as the centre of a circle; a still point of
tranquillity which does not itself move, yet is inextricably linked to each moving point on the circle’s
circumference. ‘Time,’ Plato observes, ‘is the moving image of eternity’ (360BC [1998], 37E6–38A6), and
thus without a beginning or an end.

While this divine stasis was typically considered to be full of potential and creativity, ordinary time was C1.P5
perceived as profane and repetitive, with many surviving accounts suggesting that everyday time was
nothing more than a monotonous cycle of expectation of events, enacting of those events, and being aware
of events that have already occurred. In order to regulate the individual’s biological lifespan, timelessness
was apportioned into smaller units through the cyclical process of events and activities. For many early
societies, the primary way of a ording meaning to those seemingly trapped in these mundane rotations was
to imbue it with the sacred, through festive or religious events. Music in ancient times was almost
exclusively used during these events, to help create re ective environments outside the everyday where a
connection to sacred stasis might enable daily life to become spiritualized. In premodern South Asian
cultures, for example, each musical note was considered to be a manifestation of a di erent celestial being
who could be invoked only through recitation of the ‘holy melodies’ around which many ancient texts were
written.

In the West, similar connections between music and divine nature were drawn in ancient Greece, where C1.P6
music’s two main functions were in religious rituals or drama. In his treatise on poetics, for example,
Dionysius of Halicarnassus notes that only the natural metric quality of syllables should be employed in
music, to ensure a temporal ‘consonance’ or synchronization that is in line with the harmony of the cosmos.
He notes in particular that chronoi (basic units of time) must not be regulated by syllables, but rather the
syllables by the chronoi. In recitations of theatrical works by writers like Euripides or Aeschylus, these
natural rhythmic schemata were intended enable space for an audience to best re ect on the moral and
philosophical terms of the works—although, it is worth noting, rather than creating an atmosphere of
p. 3 stillness and contemplation, the natural rhythmic patterns created the writers’ lyric verse would often be
of considerable intricacy. When heard today, such surprisingly complex music o ers a powerful reminder of
just how overlooked the development of rhythmic and temporal thought has been in musicology.

By the Middle Ages, there was an even clearer connection between the expanse of mythical time and the C1.P7
human process of division. While art still depicted time (or rather timelessness) as an ‘ideal’ state to which

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
only God has access—for example, in the expansive polyphony of the Notre Dame school, often considered
to be a musical portrayal of the Boethian notions of eternity—it was increasingly the making of human time
that became centre stage. The singing of Gregorian chant bridged a gap from ancient models of music as
linguistically based, through the patterned succession of long and short values, to a more owing, concurrit
style which prioritized simplicity of shape over rhythmic complexity. As this developed into the falsobordone
style of the fteenth century, the practice gradually died out, though even into the seventeenth century,
composers such as Monteverdi and Caccini included musical approximations of the natural speech rhythms
1
in an attempt to maintain elements of this stile rappresentativo (or stile parlando) in performance.

The continuation of the church’s disciplining of the day into canonical hours (known as Divine O ce) from C1.P8
several centuries earlier provided a fertile environment for the advent of modal rhythm—a regulatory
patterning of long and short durations into a taxonomy of ligatures—towards the end of the twelfth
century. Just like the ecclesiastical regulation of time which preceded it, the systematic division of the
perfect long into shorter values mirrored the partition of God’s ‘perfect’ and uni ed timelessness into
imperfect—but practical—worldly increments. Combined with the widespread systemization of musical
2
notation through several seminal treatises, this categorization a orded musicians the tools to create more
rhythmic density through an increased use of polyrhythms and other intricate patterns.

The rupture entailed by this system of mensuration was unprecedented, and often caused practical C1.P9
problems for less skilled musicians. As a way of responding to this new demand for metricity, a gradual
externalization of temporal structuring began to take place in the form of a beaten tactus. This basic unit of
measurement was considered equal to a heartbeat or a breath (c.60–67 bpm), and brought the relationship
3
of duration, succession, and motion into allegiance with the microrhythms of bodily movement. While the
rhythmic feel or tempo of the piece would vary depending on the conditions around it, this basic pulse
provided a unifying frame of reference which grounded rhythmic uctuations in musical expression. It is no
coincidence that such a desire for an external regulatory timekeeper occurred around the same time the rst
accurate clocks were made.

With the move towards score-based notation, rather than individualistic part-books, came an invitation to C1.P10
codify the collective experience of time even further. This shift from proportional mensural procedures to
an orthochronic system brought with it many features of modern-day temporal notation, including beams,
p. 4 ties, and bar lines. The de ning feature of this new notation was its xed relationship to each symbol,
whose temporal values remained constant regardless of context. As part of this regulatory process—one
which Margaret Bent (2002) observes was distinctive in this period in being driven by shifting performance
practices rather than church ideals—came the quantifying of tempo terminology around the turn of the
seventeenth century into a set of standardizations known as tempo giusto (literally, ‘just’ or ‘suitable’ time).
The idea of a mutually agreed speed for a piece of music had been implied for years in treatises on the tactus,
but the implication that knowing such a speed was now part of the basic toolkit for a working musician was
in many ways a radical notion—one which inevitably received a lot of attention from sixteenth- and
seventeenth-century music theorists.

The practical considerations of this new notation system invited musicians and composers to be more C1.P11
overtly aware of metricity in their music making. With regularity came hierarchy, and the unlocking of
deeper rhythmic structures and forms of metrical tension and release (from a basic anacrusis through to the
most complex of hemiola). This impetus to explore temporality as a central and exible tenet of musical
expression also led to new experiments in the ‘middle’ temporal elements of structure and narrative, which
John Butt (2010) and others have suggested was a new form of compositional and expressive virtuosity.
Even in the contained, simple forms of dance suites and other courtly entertainment—the increasing
popularity of which demonstrated another practical reason for the advent of tempo giusto—composers were
nding ways through the logic of imitative counterpoint, with its multiplicity of paradigmatic
permutations, to play with an audience’s temporal expectations. Baroque musical invention o ers a unique

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
simultaneity of experience: both statically ‘in the moment’ with its long phrase-lengths and consistent
rhythm groupings, yet highly directional and teleological as voices and harmonies are traced through their
journey towards resolution. These tensions can perhaps best be understood as two types of temporal
experience: the metric and the perceived momentum of time. As the theorist Bénigne de Bacilly observes:

[There] is a certain quality that gives soul to the song, […] called Mouvement because it stirs up, I C1.P12
may say it excites, the listeners’ attention, in the same way as do those who are the most rebellious
in harmony […] it inspires in hearts such passion as the singer wishes to create, principally that of
tenderness […] I don’t doubt at all that the variety of Mesure [number or quality of beats], whether
quick or slow, contributes a great deal to the expression of the song. But there is certainly another
quality, more re ned and more spiritual, that always holds the listener attentive and ensures that
the song is less tedious.

(Cited in Scheibert 1986: 40–41)

There is something compelling about the way in which the baroque idiom invokes a speci c process of C1.P13
expectation and memory in order to recon gure a steady ow of music, at least on the surface, into
something that appears to have multiple ways of persisting in time. Temporality during this period is not
the problematic and weighty concept that it becomes for later generations, but rather a space for musicians
p. 5 to explore freely without being tied down. Wilhelm Seidel suggests that ‘[baroque] music, while it is
going on, leads to a forgetting of time. Time is not music’s object: music neither grasps, alters, nor
interprets time. It has time only as a space for observation’ (cited in Hasty 1997: 25). Susan McClary points
to a lack of political urgency experienced by the upper classes in this period, leading to an eroded sense of
political agency that allowed intellectuals and aristocrats the luxury of time to ponder and dwell in an ‘ideal
of attentive motionlessness’ (Lucien Goldmann, cited in McClary 2012: 255). As McClary infers, there is
something rather decadent happening within music that mirrors the courtly practices that much of it would
have accompanied; its meandering harmonies, lazy inégales, and amboyant surface ornamentation
underscoring the whiling away of hours by a decadent aristocratic class. ‘[W]e could explain the music as
the pragmatic means to an autocratic end,’ she notes, ‘[with] the deliberate anaesthetizing of a potentially
restless group of subjects. The geometrical gardens at Versailles, the carefully executed divertissements, the
highly regulated dance maneuvers […] were deliberately designed to produce these political e ects, to lull
aristocrats into habits of activity- lled oblivion’ (2012: 254).

We end this brief survey of developments and junctures in musical time at the point where the volume, with C1.P14
its emphasis on contemporary understanding and experience of musical time, begins.
Overview of the Volume C1.S2

The book is organized into four parts. Part I, ‘Framing Musical Time’, contains a set of critical approaches to C1.P15
musical time that address questions of meaning and experience informed by philosophical concerns. Part II,
‘Cognition, Action, and Experience’, moves from philosophical critique to the models and mechanisms that
underpin temporal experience. Part III, ‘Metrics and Temporal Organization’, addresses questions of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
timekeeping and the construction of musical beats and metre, as well as instances of human–machine
temporal relations. Part IV, ‘Cultures of Time’, examines some of the radically di erent animations of
musical time in distinct genres or traditions.

The chapters here represent a range of disciplinary questionings, and within the diversity of approaches in C1.P16
the writing, we can see many points of intersection and dialogue across the di erent parts of the book.
Running through these broad perspectives are questions and debates that have underpinned much of the
scholarship on time and temporality in music: the paradoxical relationship between time felt as a
continuous stream of experience and yet charged with a feeling of discontinuity; musical time as a
social/aesthetic construct that yet o ers transcendence, seeming to lie beyond our making of it; musical
time’s capacity to shape the immediacy of intersubjective relations and sociality; musical time as tense—
what has gone, what is about to be, and what is now—which relates to questions of memory and
association, introspection and expectation, and immediate experience.

p. 6 Framing Musical Time C1.S3

Part I of the volume includes writings that bring philosophy and music into dialogue, as well as chapters C1.P17
that, while less philosophically driven, do important work in ‘framing’ or contextualizing musical time and
our experience of it. The questions around time and musical time in philosophy are abundant. While much
Anglophone philosophy on time—for example, J. M. E. McTaggart’s (1927) work on the (un)reality of time,
expressed through the A- and B-series—depends on clarifying the nature of tense and tenselessness, the
dependency or otherwise of truth relations to time, and the signi cance of subjectivity in relation to the
time of the world, has tended to grant primacy to the B-series perspective, much continental philosophy
has expressed concern for time in thoroughly subjective, A-series terms. The nature of music as an art
practice seems to demand this, and much of the work in contemporary philosophy of music develops
through this concern for experience, perhaps most obviously in the work of Bergson and Husserl, and then
latterly in the work of Deleuze, whose interest in music is made use of in some of these opening chapters.

Andrew Bowie (Chapter 2) opens this rst section with a wide-ranging assessment of musical time and C1.P18
philosophy in the modern world, picking up some of the themes brought out in the rst half of this
introduction. The nature of thought itself in modernity underpins much of our changed understandings of
time, and the chapter opens with some re ections on the greater signi cance of temporality in music of the
modern period. Music is seen to enact many of the meanings of time that had prior signi cance in religious
thought or philosophy and also, in common with other art forms, overcomes the ‘futility and inevitability’
that characterizes some of the radical understandings of time associated with the physical sciences. Key for
the purposes of this book is the chapter’s examination of identity and di erence; Bowie draws on Kant to
characterize rhythm as exemplifying this problem of experiencing music and its temporal constituents (i.e.
rhythms) as distinct; separated but only understandable in their being part of a whole. Rhythm therefore
becomes a key mechanism for meaning—for making sense and being able to live in the world. Part of this
sense-making lies in the ordering capacity of rhythm, and drawing on Schegel, Bowie cites the ‘role of
rhythm in calming young infants and them the sense of a structured world’, an idea that chimes with
Colwyn Trevarthen’s (1980) work on the rhythmic basis of our intersubjectivity. It is within phenomenology
that Bowie sees the greatest opportunity for moving beyond an observational stance towards objectifying
music and time, towards a participatory stance, by which he means drawing on the lived experience of
practice as opposed to relying on theoretical insight. Bowie concludes, not by dismissing theories of musical
and temporal perception, but by encouraging greater attention to the phenomenological, and arguing that
by engaging with a philosophy of music that is emergent from music, rather than making music its object,
we get closer to an understanding of time, music, and their meanings.

p. 7 In a chapter that brings examples of nineteenth-century music into temporal reckoning, Lawrence Kramer C1.P19

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
(Chapter 3) looks at the animation of musical and sociohistorical time during a period of intense, and
sometimes contradictory change. Like Bowie, Kramer is concerned with shifts in temporal meaning in the
period leading up to the modern world, but looks to three temporal modes to frame his argument. Epochal
time (large-scale cosmological time), novelistic time (layered and complex), and railway time (as the lived
experience of the railway journey) are seen to correlate with symphonic expansiveness, and the multi-
movement design within symphonies and, at the other end of the scale, the development of the
miniaturization of pieces, particularly for piano. A compelling feature of this expansiveness that stretches
between the novel and symphonic form is the sense of absorption common to both; Kramer cites Schubert’s
Symphony No. 9 in C major, D 944 (1825–8) and Schumann’s Symphony No. 9 in D minor Op. 120 (1841, rev.
1851) as distinct examples of novelistic time—in the Schubert, being aware of the novel is seen not as a way
into creating a narrative (a programme) but as an analogue of the process of the telling; whereas in the
Schumann, a novelistic temporality is made much more explicit. Epochal time—a concept that developed in
the late eighteenth century—by contrast, informs symphonic work of that period through the spread of a
preoccupation with the unimaginably long periods of time associated with evolution and geological time. A
number of works draw on this ampli cation of time, but Kramer points to César Franck’s Symphony in D
minor (1888) in its use of deep time, a slow evolution that also points to cosmological time’s beginning and
ending. The chapter ends by looking at changes in transportation that shrank distance and required greater
precision in the measurement of time—namely, the spread of railways. Kramer develops a reading of
Chopin’s Preludes Op. 28 (1835–9) and their homological compression of time, not only a move that
re ected transportation but also carried an accompanying sense of time being domesticated and more
private.

In the following two chapters, Kristina Knowles (Chapter 4) and Anne Danielsen (Chapter 5) take rather C1.P20
di erent routes to explore the relationships between our experience of time and the musical materials that
may animate such experience. Knowles suggests the possibility of music’s timelessness—a possibility that
appears to belie the complementarity of worldly time and musical time. Knowles focuses on the idea of
timelessness in contrast to a more objectivist idea of music and time, in which both are subject to external
validation and measurement. Central to Knowles’s position is the part of the listener in making sense of
musical time, and she turns to Raymond Monelle’s distinction between lyric and progressive time as a way
to map the di erent temporalities that more static (lyric) or more dynamic (progressive) musical passages
convey. This allows her to move towards the greater importance of listeners’ internal time perception and
how music may impact on that. Timelessness, however, seems to demand much more of a listener than the
idea that music may suggest slowing down or speeding up, and Knowles suggests it is in the idea of musical
time exceeding a temporal norm where the capacity to feel timelessness in music may begin. It is here that
p. 8 our understanding of two psychological mechanisms—the estimation of duration related to the content
of a temporal experience, and our nite attentional resources—are brought into dialogue. Both of these, in
distinct ways, may be considerably changed by musical processes (metred or unmetred music, tonal stasis,
repetition or not, and so on) and lead to a sense of time passing being thoroughly attenuated or activated.

It is the timing of funk and African American music that leads Anne Danielsen (Chapter 5) to a consideration C1.P21
of the experience of musical time. Danielsen, like Knowles, draws on ideas of repetition for understanding
identity and temporality in music. She rst provides a detailed examination of structural tension in funk by
looking at the ways in which layers of rhythm produce counterweights to one another—the movement
between tension and resolution (‘a stable instability’) that helps keep the listener engaged in the groove.
Although groove in Danielsen’s schematic necessarily involves repetition, this does not mean a simple
ostinato; rather, repetition involves the repeating of the dynamic shifts within a rhythmic pattern—the
‘changing same’ that Amiri Baraka applied to jazz tradition, writ small. In the second half of the chapter,
Danielsen moves from the analysis of funk rhythms to how they are experienced, and in doing so addresses
the fracture in moving from momentary time as experienced to time as understood. Here, she draws on
phenomenology and contemporary hermeneutics to address the issue of presence, and the relationship

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
between the pure experience of being in the groove, and our description of that; in other words, the shift to
meaning. We can see links here to the temporal switching that Anthony Gritten (Chapter 8) refers to, and
although Danielsen is not describing attention and distraction, there are commonalities between the
chapters in their understanding of the heightening of sensation—intensity as Danielsen describes it—as it
is a ected by breaks in the ow of the music and our relation to the momentary feeling of groove.

The politics of musical time is the focus for Chris Stover in Chapter 6. His writing builds through an analysis C1.P22
of the ways in which musical time may function, using ideas that radically expand the notion of subjectivity.
Musical time is seen in this account to work as a special case, and Stover argues that music can take place
without the presence of musical time, which takes on a precise meaning in the chapter. He initially looks to
the idea of the body as a means to recon gure agency and subjectivity. Bodies may be performing bodies,
listening bodies, or musical objects. In a manner not dissimilar to Latourian ideas of the non-human
actants, musical sound is seen to here to be an equal participant in the constitution of musical time; not only
that, but such bodies are connected performatively and relationally. In the second part of the chapter, Stover
turns in more detail to Rancière’s understanding of the political as a way to penetrate what it means to
create the musical meanings that might accrue to musical time. Rancière’s ideas are brought to bear
through the lens of Deleuze’s reading of time—a set of syntheses that fold past, present, and future into the
creation of musical meaning. The temporal unfolding of any musical event is then read through Rancière’s
understanding of how musical time works within the social order or logic. Within this framing, even the
expressive work of timing can be seen to exercise musical time through its disruption of a given temporal
logic, of what might be seen as a ‘proper’ reading of a musical event.

p. 9 Nathan Mercieca (Chapter 7) develops a not dissimilar theme to Stover, taking a Deleuzian perspective on C1.P23
how music might be performed in relation to the pre-ordained logic of the work-concept, exploring the
con ict that emerges from our understanding of the relationship of a score to its performance. More
precisely, Mercieca views Nicholas Cook’s writings on the turn to performance as a productive problematic
—or at least as an invitation to problematize the ways in which we might read performance. Two distinct
temporalities are seen to demarcate the scope of a performance: the rst is the ‘transhistorical’ notion of
relations between performances being key to our making sense of any individual performance (an
intertextuality), and the second is predicated on the sense of a local, single performance having a certain
autonomy, or at least being self-supporting. This tension between the two begs the question of how a work
maintains ‘a relationship to its own past’—indeed, whether we need to consider such a position. Mercieca
uses temporality to explore the work-concept here. In doing so, he develops a reading of Deleuze’s
understanding of the musical work, partly in dialogue with Michael Gallope’s commentary on Deleuze and
music. Mercieca’s own argument hinges on the example of a contemporary jazz group reworking
‘September’ by Earth, Wind & Fire, and in doing so he interrogates the ways in which a piece can be
understood to have a present identity of its own. Mercieca brings some resolution to the question of how a
musical work might be formulated—not as either a performed act or as a transcendent work but in a
constant dialectic.

Part I closes with Anthony Gritten’s Chapter 8, which forms something of a bridge between philosophically C1.P24
and psychologically informed work on musical experience. Gritten shares, with Knowles, a concern for the
attentional resources that we bring to music but argues that much more should be focused on attention’s
absence—distraction—as part of musical practice, suggesting that distraction can be understood in a much
more productive sense than our natural attitude would usually allow. At the heart of Gritten’s critique is to
question the cultural politics of attending—how we are to understand the tension between the open ow of
thought and our inability to focus as part of a world that demands increasing cognitive and also creative
capital—and he applies his questions through an exploration of performers and their ways of listening,
presenting a series of phenomenological reductions centred on distraction. In arguing for a rehabilitation of
distraction, Gritten sees it as a type of productive temporal switching rather than an aesthetic problem. The

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
switch that is described sets up the relationship between attention and distraction as not opposites but
necessary complements within the dynamic of listening. Gritten, however, is going beyond a psychological
process here, and points to the need to open up the regime of listening which dominates much
contemporary musical practice by acknowledging that ‘performing takes place in the world’, and that
denying our distractions is to close our minds and ears.

Cognition, Action, and Experience C1.S4

In Part II, attention turns to psychologically informed work on musical time, or research that is derived at C1.P25
p. 10 some level from scienti c thought about musical temporality. This part comprises both empirical studies
and more theoretically informed work, and presents a number of broad themes. First is a concern for the
analysis and understanding of musical temporality across di erent scales and how to make sense of the
relationships between them—for example, the relations between momentary sensorimotor responses and
their development over evolutionary time, as Juan Loaiza (Chapter 10) indicates. Second is the question of
how temporal experience is perceived across modalities, an increasingly demanding issue in music
psychology. Third is the role of more recent work within embodied/enactive theory in accounting for the
production and perception of musical time. Fourth are questions about the universality of musical time
perception as opposed to the contingency and cultural speci city that is suggested, for example, by Rainer
Polak’s chapter (Chapter 14).

John C. Bispham (Chapter 9) opens Part II by turning to the foundations of time perception as part of C1.P26
music’s design features as they have been shaped over our evolutionary past. He articulates the idea of the
musical moment—a characterization of the ‘extended experience in action and time’ that music represents
for humans. For Bispham, this experience centres on two musical features, pulse and tone, but is also given
dynamism and meaning through the character of human motivation—part of which is the impulse towards
social connection and the ful lments of human life grounded in emotional experience. Bispham takes us
through the psychological underpinnings of time that ground musical experience, but contends that music
must be seen as related to but distinct from our general capacity to be in time. The psychological processing
of time depends on di erent neural mechanisms, which includes duration-based or beat-based strategies
to derive structure from music, and phase and period correction, which we use to adjust to beats of others.
The second of these latter mechanisms that appears to be a uniquely human corrective and one that is
intertwined with intention, conscious behaviours, and intersubjective being. While music is constituted in
timing mechanisms that can be found across humankind and blend with our more general capacities for
keeping and organizing time, this suite of mechanisms is ampli ed considerably in music: the timing of
music is crucially geared towards ‘communicative intent’ and in the idea of a musical moment provides an
extraordinary palette of ordered expression.

One of the questions that faces musicologists is how to connect the analytic categories that underpin our C1.P27
understanding of musical time. In a chapter that draws on dynamic systems thinking, Juan M. Loaiza
(Chapter 10) lays out a non-reductionist theoretical framework which interweaves the di erent layers of
temporal scale from which musical time emerges. Loaiza builds on the set of timescales through which we
understand duration, from the nest subatomic scale through to cosmological time, yet identi es a
tendency to see the relationship between these scales as inescapably reductionist—ad extremis, this can
lead to something like the grooviness of a samba being reduced to the observation of neuronal activity. Part
of his response is to recast time scales as ranges within which we can see nested, interacting temporal
relationships at work. Slower and faster timescales, relevant to musicking, are therefore seen in his model
as thoroughly interdependent and extended across sensorimotor activity, social systems, and the life cycle
p. 11 of the self. Each of these can be seen as a set of ‘spans’—for example, the sensorimotor range brings both
the very small timescale of sensorimotor adjustment and the massively slow timescale of species evolution
into focus as a relationship between body and environment. The second range involves social interactive

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
timescales and historical processes, and the third range is the continuity between a person’s life history and
their sense-making. Each of these very di erent forms of temporal range is itself connected to the others
through what Loaiza terms ‘transitions’, the ways in which processes in one temporal range become ‘re-
organized and re-calibrated’ in the next. Both of these transition mechanisms can be viewed as processes
involving the reduction of divergence over time and between social interactants.

Rolf Inge Godøy (Chapter 11) provides a more insistent focus on the smaller scales of musical perception, C1.P28
and addresses one of the paradoxical problems in understanding sense-making in music—the perceptual
relationship between the discrete events of music and the sense of continuity that emerges out of these. He
uses the idea of musical instants to describe the fragments of sound that not only capture the fundamental
character of a sound but may also o er a more complex aesthetic, a ective picture that registers through
our associating such fragments with images of sound/body motion in which an instant may be grounded.
Sound events are composed of the complex interaction between the features of the sound and our picking up
on their ecological signi cance, and this leads Godøy to include the close relationship between body motion
and a sonic object as being vital to our perception of a sound. The key concept that drives the idea of a
musical instant is that of intermittency—a feature of both theories of time perception and in human motor
control, in tandem with the cross-modal idea of shapes. Intermittency describes the process whereby body
motion involves a form of chunking, discrete programmable moves which allow for a degree of open-loop
control and seems to o er more economic consumption of energy. The idea of shape to represent sonic
features has attracted considerable research: the signi cance of shape in music lies in its ability to capture
what is temporal and unfolding in a single, atemporal image; the gestalt quality of a shape as well as their
generic nature suggests that they can operate as mental constructs across di erent modalities. Godøy
brings intermittency and shape together in his notion of a sound–motion object, a multimodal gestalt that
intertwines human movement with sound, and is described in musical time as a musical instant—a means
to gain further understanding of the relationship between continuity and discontinuity in our experience of
music.

While Godøy addresses the accumulation of fragments of experience across modalities, the next chapter C1.P29
also addresses multimodality but directed towards tempo and timing. Renee Timmers (Chapter 12) brings
together two distinct areas of research: timing and movement in music, and cross-modal mappings in
relation to musical time. The movements of musicians do not simply produce timings in music but may
in uence the perception of time on the part of observers; movements a ect the perceived expressivity of a
performance, or the perceived durations of tones. Timmers argues that the nature of correspondences
across modalities—that is, how visual and aural features are seen as belonging together (e.g. large size
associated with low pitch)—is vital to our understanding of the experience of timing and tempo, and our
p. 12 ability to coordinate performance. But the picture is complex. Recent research shows that beyond certain
broad principles, the relationship between the mapping of associated sonic properties (loud and fast) can in
certain situations become less sure, and the periodic congruence of movement and sound is also more
variable than might be imagined. Timmers reports ndings from a study that explores how visual
information contributes to the capacity of a performer to synchronize with their co-performer. She
concludes by suggesting a number of points that arise out of existing work yet remain in need of more
experimental study: that actions play a central role in cross-modal correspondence—a ground for the deep
mappings that span auditory and visual modalities—and that cross-modally expressible properties in
music have a ‘prioritized’ or intensi ed meaning on the basis of their widening perceptual access to an
event.

Other chapters in the book develop perspectives on time that point to ontological perspectives, blurring the C1.P30
distinction between human and non-human actants. In exploring the beatmatching skills of DJs, Maria
Witek (Chapter 13) brings a very di erent, enactivist approach to bear on the human/machine interface.
Rather than atten the agency between human and device, Witek uses the work of the DJ as a vivid example

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
of the human skill of sensorimotor synchronization, and applies an enactivist perspective to the DJ’s art of
beatmatching. The starting point for the chapter is entrainment theory, which is used across a range of
disciplines to account for the relatively stable coupling or synchrony that can occur in mechanical, natural,
and human systems. When applied to human temporal behaviours such as music-making, one of the tenets
of entrainment theory is that much of the work to stay aligned with another occurs at an unconscious level;
this does not deny intentionality in musical expression, but speaks to the largely automaticized process of
shifting note onsets to accommodate another. Witek suggests that beatmatching operates in a
fundamentally di erent way from conventional contexts of music-making because the process of syncing is
di erent in kind, requiring an unusual degree of conscious control. In this cognitive study, Witek also makes
some important cultural points about how di erent types of musical work are treated very di erently in
terms of their signalling virtuosity and skill. The uncovering of the beatmatching skills points not only to
enactivism as a perspective on performance but also to the need to re ne our understanding of the DJ as
timekeeper.

Based on his many years’ work with musicians from Mali, Rainer Polak explores the role of uneven C1.P31
subdivisions as temporal reference structures in Malian music (Chapter 14). Polak suggests, in light of his
ndings that metres may be based on non-isochronous subdivisions, that there is a need to reframe our
understandings of basic rhythm perception away from the universal and towards the culturally speci c. The
orthodox position on our perception of basic rhythmic and metrical structures is tied to the privileging of
isochrony—that is, viewing the pickup of such musical structures as dependent on broadly equal durations
(or simple integer ratios such as 2:1) between note onsets—a perceptual starting point from which we are
able to categorize and infer temporal organization in music. These sorts of fundamental timing relations are
viewed both as biological preconditions—that is, we have innate dispositions towards picking up isochrony
p. 13 in music and in our environment—and as psychologically plausible, from the point of view of categorical
rhythm perception. Polak’s research points to a 4:3 ratio of long–short durations that constitutes a distinct,
stable rhythmic category rather than simply expressive deviation. This feature, in terms both of
enculturated listening and of musicians’ performance, suggests that isochrony—the sine qua non of
research into rhythm perception—is a perceptual bias among Western listeners, just as much as the Malian
ability to feel and make music with the 4:3 subdivision. This nal chapter of the section, with its strong
perceptual, psychologically grounded focus on metre, takes us into the third part of the book.
Metrics and Temporal Organization C1.S5

At this point in the volume, the emphasis in the contributions turns towards musical time as measurement C1.P32
and structure, both in terms of cultural practices and analytic models. The point of contact between the
various approaches taken in Part III is the strong focus on temporal structures, particularly metre; but the
other common theme between these chapters is how various forms of con ict, either perceptual or
sociocultural, are played out within musical structures and in the cultural histories of those temporal

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
structures (as shown in Jonathan Still’s Chapter 20) on dancers’ and musicians’ contrasting approaches to
metre). This con ict extends to the use of technologies in policing musical time and metre (metronomes
and digital quantizing), through to the negotiation of metred forms in performance and in recording. In
performance, this includes the work that musicians need to do to rescue a performance that is going awry
through metrical disagreement between the players—an imperative underpinned by the strong tendency to
entrain. In recording, and as shown by Andrew Bowsher’s study of early blues practice (Chapter 18) and
Landon Morrison’s work (Chapter 17), the homogenization of much performance speaks to the con icted
politics of metre and the impact of recording technologies.

The rst chapter of Part III explores how temporal features impact on musical experience through metre C1.P33
and its displacement. Mark Gotham (Chapter 15) presents a model of dissonance that quanti es the
contribution of metre in creating musical tension. Gotham’s model, building broadly on the work of Harald
Krebs, applies mathematical thought to the patterns of consonance and dissonance in the metrical
structures of a piece. At the point where strong and weak patterns in a metrical hierarchy converge with the
harmonic structure, the architectonics of a piece are consonant. When metrical accents are placed which
undermine this by providing points at odds with the prevailing structure, this can be described as metrical
dissonance, and has implications for the degree of certainty/uncertainty that a listener may feel. In setting
out a formalization of how metric dissonance may be understood in terms of types and intensity, Gotham
applies the model to Bach’s Brandenburg Concerto No. 6 in B♭ major BWV 1051 (I) (1721). Building on this
example, Gotham successively re nes a model, acknowledging that the earlier version of the model as
applied gives (for example) equal weight to the di erent metrical levels in a hierarchy in a way that serves
p. 14 the psychology of listening less well. In suggesting a number of re nements to the model, Gotham goes
on to consider mixed metres and ends with the suggestion that simple binary metres are more destabilized
by metrical displacement than mixed, ‘already unstable’ metres.

Contemporary scholars may not be aware of Johann Maelzel’s place in the temporal turn towards measure C1.P34
and regulation yet Maelzel’s metronome, created at the beginning of the nineteenth century, marks a
technological shift that had radical implications for contemporary musical ideas about tempo and
timekeeping. Alexander E. Bonus (Chapter 16) takes Maelzel’s invention as a starting point for a critical
history of timing and precision. One of the fascinations of this ‘metronomic turn’ is the anxiety that
contemporaries of Maelzel experienced with his temporal automata and their ‘automatic idiocy’. The
central component of this anxiety lay in a territory familiar to social being in the late twentieth and early
twenty- rst centuries—that of artistic expression and, more importantly, human agency in relation to
technology. In this regard, Maelzel’s works were seen as a vision of what could be: an industrialized
workforce moving in step with machinery. This contrast between the sense of automation for labour as
progressive, and automation in temporal art as repressive, is matched by the contrasting way in which
Maelzel’s ‘labour’ products failed to ride the wave of industrial momentum yet his metronomes found their
way into modern pedagogy, becoming ‘virtuous models for musical functioning’ even up to the present day.
Bonus summarizes that there is a lack of historical understanding that underpins much contemporary
investment in the idea of metronomic time; he concludes that the temporal norms of musical life in the
early twenty- rst century represent a misplaced desire to regulate human motion and achieve an ultimately
ctional temporal perfection.
Rhythm quantization can in many ways be seen as the natural successor to the mechanical metronome C1.P35
developed by Maelzel. In Chapter 17, Landon Morrison outlines the development of digital quantization,
examining its use in popular music production but situating its impact well beyond this by exploring its
reception and contribution to signi cant changes in contemporary social and aesthetic practices. The
quantization of musical timing has developed within a much larger reframing of cultural practices around
music and time. Morrison traces the move to this temporal technoculture beyond the development of
machines—instead, he pre gures the momentum towards the mechanical with the move away from bodily

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
motion as the arbiter of the beat towards a more objective conception of time in music that mirrored
Newton’s work on physics and time. In this way, Morrison sees contemporary musical time as one part of a
succession of temporal regimes ‘grounded in beliefs about rhythm and metre’: rather than its exemplifying
a digitized rupture from the human, it is just another example of the ‘externalization of social and cognitive
processes’—more an example of extended cognition, and an extension of human capacity rather than its
alienation. The contemporary use of quantization remains, however, grounded in a discourse that views
music as something that must originate from the human body. Morrison here, helpfully points to genre as
central to understanding the aesthetics of machine time. He concludes in part by making the important
p. 15 point, however, that the battle is all but over—most forms of popular music use the tools of quantization
to some degree, and it has become part of the natural attitude towards producing and listening to music.

The politics of temporal measure and its representation are further discussed in Andrew Bowsher’s work on C1.P36
African American blues (Chapter 18), which examines the racialized narratives and representation of
country blues practices from the inception of recording to the present. As other chapters in Part II show,
what might be seen as a relatively neutral space in musical discourse—time and timing—rather than simply
being part of technique in music become a complex political domain, open to the ‘privilege of domination’,
as Bowsher suggests, quoting from Fred Myers. Bowsher looks at a series of blues artists and the temporal
features of their work as the blues became recast over time to t a narrative of Black authenticity. He
questions the purity of country blues in a number of ways: the early artists played many forms of music
alongside blues numbers, and many of these artists were professionalized and highly successful, an image
far from the idea of these musicians as ‘unhindered by the trappings of commercial culture’. Signi cantly,
the timing and structure of blues songs in live performance was at odds with the contemporary technology
of 78 rpm recordings. Whereas much of the output of country blues singers celebrated their capacity to work
in a loose framework (adding or taking away whole and half measures), the format of the 78 rpm record led
to a diminishing of the exibility in country blues artists’ use of form and towards a more consistent
structure. While in some senses the early recordings do powerfully suggest an authenticity, in fact the
constraints of this technology limited how country blues would be heard away from the usual spaces of
performance. The legacy of the early country blues artists reminds us that the racialized narratives, ideas of
authentic Black expression, and the very structure of the blues are to be held up for question and scrutiny.

From blues to jazz, and from the shifting of metrical proportions over time to the negotiation of metrical C1.P37
congruence in the moment. Thelonious Monk and his quartet are the focus for Ryan D. W. Bruce’s
examination of metric displacement in a recorded performance of ‘Evidence’ by the group in 1987 (Chapter
19). The displacement at work in this recording stems not from any compositional device—although
‘Evidence’ is metrically ambiguous in Monk’s use of quarter notes throughout the head—but through the
sense of the metre becoming dislocated between members of the group during the statement of this
rhythmically complex head. The metrical dis-attunement between the group persists into the saxophone
solo that follows the head, and Bruce explores the negotiations between the players in their coming back to
metric alignment. Bruce’s chapter therefore explores displacement that occurs as a result of musicians
being attuned mistakenly to the temporal structure of a composition. The analysis that Bruce presents is a
detailed examination of the various strategies that the performers use to orient themselves towards a
metrically coherent improvisation. The di erent instrumentalists employ distinct musical signals to one
another in attempting to resolve the ambiguity of the performance of metre, and Bruce establishes that the
performance of metre and metric displacement becomes just as much a matter of improvisational dexterity
as in any conventional performance.

p. 16 Feeling the shared structures of music may not only vary between musicians within a single performance C1.P38
(as in the Monk case study above) but are clearly di erent for musicians working within distinct genres. It
may only be when we are in contact with players from di erent musical worlds that our natural attitude to
musical frameworks may be jarred or compromised. These circumstances are brought into sharp relief in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
Jonathan Still’s exploration of the ways in which metre is used by musicians and dancers as a regulatory,
supposedly common, framework (Chapter 20). The counting that dancers use in their work enjoys a
problematic relationship with musical metre as understood by musicians. This raises, Still argues, questions
that go well beyond the counting itself to the relationship between the bodily experience of the dancer and
the representational, highly conceptualized notion of musical metre. Key to Still’s exploration of dance
practice and pedagogy is non-representational theory and the notion of the ‘everyday’, ideas which frame
the signi cance of bodily practice as part of the sense-making of music and musical metre. He draws on
Rothstein’s work on Franco-Italian and German approaches to the placement of bar lines, which can be
traced to the di erent prosodic features of these languages. What becomes clear is that a simple opposition
between musical and choreographic counting is unsustainable; in fact, even in relatively simple Western
musical examples, Still shows that there is no single music-theoretic approach to metre, and that the
teaching of metrical principles according to any particular system becomes deeply ideological, erasing the
contingencies of cultural transmission in favour of mythic, universal principles. For Still, in the interests of
creative practice, the felt, mobile appreciation of music in a dancer must be held in a dialogic, productive
tension with the representation of metre.

Cultures of Time C1.S6

The nal section of the volume, Part IV, turns towards temporal experiences and practices that are de ned C1.P39
by, rather than simply being examples of, their cultural speci city—a concern for time and timing as lived
by di erent groups. As ethnomusicologists know well, the notion of a culture is problematic. First, there are
certain ways in which musical time has something approaching universality in human being—its capacity
to entrain, provide ritual and functional structure, form part of our emotional response to music, and so on;
and second, within a culture there are innumerable, radical di erences to be found in the way that people
experience and make sense of musical time. Nevertheless, these nal chapters highlight the utility and
necessity of studying musical time within particular cultural groupings. By emphasizing the idea of
temporal cultivation over temporal cultures, these chapters examine the practice of musical time within
scenes, groups, audiences, and for composers, and bring together more prototypical examples of a musical
culture at work alongside more di use or unusual cases of temporal making in music.

Although we tend to associate the idea of making time with performers, in Chloë Alaghband-Zadeh’s C1.P40
p. 17 Chapter 21 on North Indian classical listeners, there is a strong sense of the work that audience members
carry out in the shaping of temporal practice as part of performance. Zadeh’s work focuses on
connoisseurship among North Indian classical music lovers—or rasikas, as they are known. In paying
attention to this group, Alaghband-Zadeh addresses the need to understand musical time not only as a set
of structures or even performance processes, but as inclusive of the listener and what they bring to the
moment of performance. Drawing on scholarship on mediation, she identi es some of the problems
attached to a structural or homological reading of musical time vis-à-vis sociohistorical time, and proposes
going beyond music’s mediation of time towards a consideration of the ways in which multiple
temporalities come together ‘in speci c acts of performing and listening to music’. For the rasikas who
participated in Alaghband-Zadeh’s ethnographic study, music becomes a temporal resource for engaging in
a ‘leisurely temporality’, a temporality that is addressed in terms of its wider social context and
consequences. The time of performance for these connoisseurs is su used with a sense of times past, and
music being a world removed from modernity and its accelerations. Music is a leisurely site of resistance to
many aspects of contemporary India, drawing on multiple timescales between being in the moment of
performance and occupying a place in much longer temporal scales that span modernity.

The distinctive performance mode within Indonesian gamelan, known as palaran, is the focus of Jonathan C1.P41
Roberts’s writing in Chapter 22. Roberts explores the con ictual relationship between the tacit rules of
timing and the momentary negotiations that occupy the musicians in this virtuosic and often risky

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
performance style. The palaran form of singer and small ensemble emerges from the unaccompanied
singing tradition of macatan, and it is this translation from a relatively free interpretive mode to a more
interactive musical space that sets up the complex set of tensions that con gure the performance. Two
particular areas of performance are salient here: rst, the way in which the instrumentalists in the
ensemble respond to the singer’s delivery; and second, how the singer and instrumentalists approach
particular, strong cadential points when a gong is to be played. Both of these areas are opportunities for
excitement among players and audience alike; these represent moments of risk when connoisseurs in an
audience can pick up on the tensions between the players as they attempt to both accommodate the
exibility of the singer and adhere to the structural imperatives of gamelan, particularly in terms of the
placement of the gong note. These con icting demands of structure and exible support for the singer can
lead to a loss of coherence when the aesthetic demands smoothness and e ortlessness. Yet, in the failure to
achieve these forms of cohesion lies the excitement and popularity of the genre, and the opportunity to deal
with musical risk while allowing for virtuosic extemporization.

With a nod to some of the earlier chapters in the volume, Nathan C. Bakkum extends the moment of C1.P42
improvisation through successively increased timescales by focusing on the performances of a network of
New York-based jazz musicians (Chapter 23). The distributed creativity at work between these players is
seen as occurring over three layers of time. Bakkum describes how, in the instant of performing, there is a
p. 18 highly interactive logic at work which musicians characterize as, in part, about working for others, but also
about attempting to get to a state where there is a relatively light cognitive engagement. At a meso-level,
there is a switch towards a broader scaled attention to the compositional aspects of performance. This
shaping of ideas and approaches within performance coalesces at the level of continued work as an
ensemble and the development over time (frequently over the course of a tour) of compositions that may
become changed over a period through the process of shared creative practice. The broadest scale, at the
level of the traditions of jazz culture, is established through lineage and the shared histories of musicians.
Lineage not only involves direct connections with the tradition through teachers and more experienced
players, but also suggests the way in which musicians represent themselves through asserting their
connections to iconic or well-known players of the past. Such shared histories provide storehouses of
knowledge as well as the foundation for creative a nities. Although analytically separable, these layers of
time that make up a performance are thoroughly interpenetrating, and it is within these layers that the
musicians nd resonance ‘in each gesture and across generations’.

The role of time in social bonding is well recognized within, for example, the entrainment literature. More C1.P43
broadly, time and timing, within or outwith music performance, is seen as occupying a central role in social
interaction, whether that be in understanding turn-taking in discourse or gesture within musical
performance. Emily Payne takes John Cage’s Concert for Piano and Orchestra (1957–8) as a case study of
social interaction in conditions of temporal indeterminacy (Chapter 24). In contrast to Witek’s chapter
(Chapter 13), here the questioning of temporal engagement turns away from machine to human
performance and equally turns from cognitive control to disruption. Payne examines a performance by
Apartment House of Cage’s work, and characterizes the relationship between the temporal indeterminacy of
that work and the sociomusical interaction within the ensemble as reported by the musicians, understood
by Payne as a performance mode characterized as a ‘separate togetherness’. Cage’s work represents a shift
not only in the perceptual and productive aspects of timekeeping—none of the conventional rhythmic
structures are available to the musicians—but also through the recon guring of the role of the conductor
and the resulting social interactions between the conductor and ensemble. The study shows how di cult it
is to suppress the feeling of, and desire for, a shared temporality. Playing ‘together’ in this piece becomes
limited to starting and nishing at the same time; and yet, as Payne’s work shows, at di erent points, and
across di erent attentional levels, there are distinct forms of togetherness at work; here, the role of the
conductor, while utterly changed in some respects, still retains its catalysing powers in pushing the
ensemble in surprising, eventful ways.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
Alfred Schütz’s phenomenology of musical interaction remains an in uential account of what it is to C1.P44
perform with others. Floris Schuiling’s examination in Chapter 25 of the deployment of di erent notations
within improvisational settings is used to develop a critique of Schütz’s work on music, and also
rehabilitates Maurice Halbwachs’s productive account of temporality and sociality in music. Underpinning
Schuiling’s chapter is a dissatisfaction with a number of binaries that could be seen in terms of the
ear/improvisation—a symbol of the oral, the immediate and the authentic—posed against the eye/notation
p. 19 as a measure of consideration, intellectualism, and inauthenticity. In working through and against this
set of oppositions, Schuiling refers to his ethnographic work conducted with various improvising groups in
the Netherlands. Out of these case studies, he shows how the binaries which are often deployed do not t the
experiences of the musicians. His analysis in this chapter is that notation, far from being a technical device
—something that stands out from time and from Schütz’s perspective, outside social interaction, instead
becomes a much more signi cant element in musical sociality and in the managing of di erent
temporalities between performers.

The relation between the rhythms of modernity—its temporal logics—and the dynamics of musical C1.P45
temporality are explored in Samuel Wilson’s Chapter 26, in which he lays out the ways in which music may
mediate the time of late modernity. Wilson uses New York City in 1983, and two pieces of music from that
place and year, as a means to expose the contradictions inherent in the complex temporalities that
constitute the experience of modernity. Bill Fontana’s sound sculpture Oscillating Steel Grids Along the
Brooklyn Bridge and Morton Feldman’s String Quartet No. 2 form the basis of this critical discussion. In
Fontana’s work, the sound of the urban landscape, speci cally the sound of Brooklyn Bridge and its tra c,
was ampli ed through microphones and then relayed to another location in downtown New York. This
sound art played out and broadcast the rhythms of daily life in New York; the work called attention to the
life of the city and its history (the bridge was 100 years old in 1983) and was unrepeatable—a temporary
installation. Feldman’s work is equally transformative but works on the listener in a very di erent way
through his rejection of conventional musical time, through a number of devices which connect with some
of Jameson’s diagnostics of late capitalism—a reduction to the present, a rejection of depth and
connectedness. In summing up, Wilson, drawing on Lefebvre and Foucault, o ers an answer to how such
di erent works nevertheless respond to the conditions of lived time in the modern world.

The volume closes by inviting a further consideration of the accelerative time of the late twentieth and early C1.P46
twenty- rst centuries and the musical response to it. In a suitably contemporary tone to end the volume,
Toby Young (Chapter 27) presents an analysis of drum and bass, suggesting that this genre presents a site of
temporal experience in which the contradictions of modern temporality get played out—in part, drum and
bass acting as a palliative to the conditions of modern life as well as highlighting or intensifying the
con icted temporalities that confront people. Although the idea of accelerationism has been used to
describe the trajectory of contemporary time, it has been con gured in very di erent ways within the
scholarship. Both embracing and resisting this speeding up have been promoted as responses within the
literature, but Young radically reimagines this dialectic by arguing that drum and bass has a unique ability
to invoke multiple experiences simultaneously, which in turn provides clubgoers with an ontological
experience of revelation that illuminates the contradictions of contemporary lived time in an embodied way.
He concludes by suggesting that the plurality of experience o ered by drum and bass has the potential to
facilitate a powerful form of resistance to contemporary life through both the sonic and social experience
that the music o ers.

p. 20 The range of ideas expressed in the book, on one level, point to a set of perspectives as wide as the temporal C1.P47
imaginings that musicians and listeners have brought to music over the last 100 or so years, but also point
to the enduring concerns of humans with the time of their lives. Musical time and its passing yields, as so
many chapters in the book show, deep insight into the conditions of subjective time—time as part of our

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
interior life—which music seems to penetrate. The time of music, however, is crucially directed towards the
sense of being in time with others. Although the immediate socialities of music are con gured in
innumerable ways through the di erent elements of the musical fabric, it is time and timing that appear
most salient in music’s capacity to set up conditions of social bonding and the productive aspects of our
social relations. Going beyond the relatively immediate feeling of time as a subjective ow and as part of
human relations, the book also captures musical time as part of wider imaginings about the cultural politics
of music, and music’s increasingly important role as both mirror and hammer in shaping the temporality of
late modernity. We hope this book brings insights into and further questionings about musical time as
private and public, shaped and shaping, metric and immeasurable.

Notes
1. It is worth noting that contemporaneous performance practice slowly began to abandon this tactic as musicians C1.N1
increasingly interpreted the rapid, even notes of the style literally rather than following the natural shape of linguistic
declamation, as had been the earlier practice.

2. The most prominent of these, Ars cantus mensurabilis (ʻThe Art of Measured Songʼ) by Franco of Cologne, was notable for C1.N2
being the first publication that attempted to introduce ʻindividual, sign-specific durational valuesʼ and rudimentary time
signatures.

3. Bourdieu tellingly calls these micro-rhythms ʻtempoʼ (1977: 6–7). C1.N3


References C1.S7

Bent, M. (2002). Counterpoint, composition, and musica ficta. Routledge. C1.P50

Bourdieu, P. (1977). Outline of a theory of practice (trans. R. Nice). Cambridge University Press. C1.P52
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468821 by National Science & Technology Library user on 26 May 2023
Butt, J. (2010). Emotion in the German Lutheran baroque and the development of subjective time consciousness. Music Analysis C1.P53
29(1–3): 19–36.
Google Scholar WorldCat

Gross, D. (1985). Temporality and the modern state. Theory and Society 14(1): 53–82. C1.P57
Google Scholar WorldCat

Hasty, C. (1997). Meter as rhythm. Oxford University Press. C1.P59


Google Scholar Google Preview WorldCat COPAC

Langer, S. K. K. (1953). Feeling and form: A theory of art developed from philosophy in a new key. Charles Scribnerʼs Sons. C1.P61
Google Scholar Google Preview WorldCat COPAC

McClary, S. (2012). Desire and pleasure in seventeenth-century music. University of California Press. C1.P62
Google Scholar Google Preview WorldCat COPAC

McTaggart, J. M. E. (1927). The nature of existence, vol. 2. Cambridge University Press. C1.P63
Google Scholar Google Preview WorldCat COPAC

p. 21 Plato (360BC [1998]). Timaeus. Trans. B. Jowett. Project Gutenberg. https://www.gutenberg.org/files/1572/1572-h/1572-h.htm C1.P68
WorldCat

Scheibert, B. (1986). Jean-Henry DʼAnglebert and the seventeenth-century clavecin school. Indiana University Press. C1.P72
Google Scholar Google Preview WorldCat COPAC

Trevarthen, C. (1980). The foundations of intersubjectivity: Development of interpersonal and cooperative understanding in C1.P74
p. 22 infants. In D. Olson (ed.), The social foundations of language and thought: Essays in honor of Jerome S. Bruner, 316–342. Norton.
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
CHAPTER

2 Time in Music and Philosophy  C2

Andrew Bowie

https://doi.org/10.1093/oxfordhb/9780190947279.013.1 Pages 25–C2.P71


Published: 08 December 2021

Abstract
The kind of sense music makes, is bound up with how forms of meaning, including in verbal language,
are connected to time. Traditional cyclical, epic, and goal-oriented senses of time all play a role in
modern musical forms, even though the mythical, religious, and metaphysical content of these forms
is hollowed out by scienti c advances. The chapter considers aspects of the work of Descartes,
Rousseau, Kant, Schelling, Friedrich Schlegel, Dewey, Hegel, Merleau-Ponty, Bergson, Nietzsche, and
Heidegger, focusing particularly on the relations between self-consciousness, time, and rhythm, and
on how these contribute to the constitution of meaning. The chapter argues that it may make more
sense for philosophy to attend to what music reveals about time which can only be grasped by active
participation in music, than to seek a comprehensive explanatory account of music and time.

Keywords: time, rhythm, self-consciousness, philosophy of time, meaning, perception, science


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Science, Music, and Time C2.S1

THE ‘philosophy of time’ in the analytical philosophical tradition does not tell us a great deal about music, C2.P1
because it is largely concerned with time as an aspect of the objective universe described by the modern
natural sciences. Is time ‘real’, for example, or solely a way in which relations between objects appear to us?
Although perception of music plays a role in what such philosophy investigates, its investigation of the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
perception of music is not inherently di erent from investigation of the perception of any other
phenomenon. Discussion of music outside the domain of philosophy which predominantly orients itself in
relation to the natural sciences, on the other hand, seeks to understand music as a temporal art which o ers
unique ways of understanding time. As such, it is perhaps not surprising that the perception of music is very
often used as an example even in science-oriented accounts of perception of time. On the one hand, then,
music can be just part of the examination of how separate physical phenomena in time are perceived as
connected into something larger; on the other, the kind of sense music makes, which gives it great
importance in many people’s lives, seems intimately bound up with how forms of meaning, including in
verbal language, are connected to time. This di erence is already adumbrated in the Greek distinction
between chronos, sequential time that can be mathematically ordered, and kairos, which in Christian
theology denotes the right or opportune time for something to be said or done, so connecting time to
religious or social meaning.

This chapter will focus on philosophical issues concerning music and time in the modern period. Music in C2.P2
this period largely ceases to be regarded, for example, in Pythagorean terms, as a re ection of the harmonic
order of the cosmos, or, in Platonic terms, as a form of imitation of states of the soul that, in the name of
p. 26 preserving social order, needs to be governed by norms. In Pythagorean and Platonist conceptions, which
also dominate the Middle Ages and much of the Renaissance, music consisted of Harmonia, Rhythmos, and
Logos: ‘By Harmonia one understood regulated, rational relations of notes brought into a system, by
Rhythmos, the temporal order of music […] and by Logos, language as the expression of human reason’
(Dahlhaus 1978: 14). Ideas about music, ancient and modern, almost universally involve notions of form and
harmony. However, the signi cance of temporality in music becomes much greater in the modern period,
when the assumption that there is a ‘ready-made world’—whose pre-given, and thus timeless, order is
re ected in the mathematical proportions in music—gives way to the idea that what the world is is also a
product of human thinking and action.

Instead of thought simply representing what is always already objectively there, it now is seen as making C2.P3
things manifest in new ways that would not exist but for our interactions with the world. Before the modern
period, time tended to be seen, in many forms of mythology, as a recurring cycle that is echoed in the
seasons; in forms of epic, as a sequence in which the elements of the sequence have their sense present
within themselves, with the sequence having no ultimate goal (so involving what Erich Auerbach in Mimesis
refers to as ‘a uniformly illuminated, uniformly objective present’ (Auerbach 1953: 7)); and, in certain forms
of religion, as something whose ultimate meaning will be revealed when it reaches its goal/end. Cyclical,
epic present, and goal-oriented senses of time all play a complex, evolving role in modern musical forms,
such as rondo, variation, and sonata, even though the explicit mythical, religious, and metaphysical content
of these forms of time is in many respects hollowed out, particularly by advances in scienti c knowledge.
This suggests that such forms still play a role in how we relate to time, albeit one that can no longer be
grounded in religion or metaphysics, and so has to be enacted in what Ernst Cassirer terms ‘symbolic forms’
like music.
The Subject of Music C2.S2

One source of the hollowing out of mythical, religious, and metaphysical content is the modern shift of C2.P4
philosophical attention from the objective world to the subject associated with René Descartes’s attempt to
combat scepticism about knowledge of the external world via the certainty inherent in the subject’s
re ection on its own thinking. It is then not surprising that Descartes is one of the rst theorists to think of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
music primarily in terms of the listener. In the Compendium Musicae of 1618, he claims that music needs the
subject’s active participation if the di ering bars of a piece are to be made into a perceivable unity: ‘when we
hear the end we recall at this instant what there was at the beginning and in the rest of the song’ (Descartes
1987: 60). This idea was already adumbrated by St Augustine, whose view of time:

is based on our experience of music. When we listen to a hymn, the meaning of a sound is given by C2.P5
p. 27 the ones that come before and after it. Music can occur only in time, but if we are always in the
present moment, how is it possible to hear it? It is possible, Augustine observes, because our
consciousness is based on memory and on anticipation.

(Rovelli 2018: Kindle Locations 1604–1606)

This idea of the subject overcoming temporal separation to bring about a unity which makes something C2.P6
intelligible is crucial to the issue of time and music. The subject is itself in one sense temporal, and yet, on
the assumption that it can no longer rely on a metaphysically guaranteed order of time, it needs time to be
ordered in new ways for things to make sense. In some way, then, the subject must transcend time.

The forms in which time is comprehended therefore take on new signi cances in modernity: separation C2.P7
from traditional temporal order opens up new expressive responses. These celebrate new possibilities, but
also seek to come to terms with the loss of metaphysical certainties and with a concomitant transformation
of the sense of human nitude. Objecti ed forms of time in the sciences can lead in the direction of an
existential sense of futility and inevitability. This is epitomized by the idea of the ‘heat death of the
universe’, which sees time in terms of entropy, the irreversible increase in disorder that is inherent in the
nature of all physical change. In contrast, many of the responses to time in art in the modern period derive
from the possibilities of overcoming objective chronology and di erentiating and opening up new ways in
which time can be articulated and experienced in order to give rise to new, temporalized forms of meaning.
This will be one reason why music takes on a new status in the modern period.
Time and Space C2.S3

Jean-Jacques Rousseau, himself a composer as well as a major thinker with respect to the social and political C2.P8
demands of the modern era, sets out one aspect of what is in question here in the Essay on the Origin of
Languages: ‘The eld of music is time, of painting is space. Multiplying sounds heard simultaneously, or
developing colours one after the other, is to change their economy, it is to put the eye in the place of the ear,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
and the ear in the place of the eye’ (Rousseau n.d.: Kindle Locations 27364–27366). The perception of a
painting can admittedly share features with the perception of music—in that the eye must range over
di erent aspects of the painting and connect them to make sense of it, in ways analogous to what the ear
does in music—but the sensory manifestation of each clearly di ers. The important issue is, therefore, how
the relationship between time, which is seen as moving, and space, which is not, is interpreted. When
‘movement’ is used in relation to music, it does not refer to objective movement of the kind measured by
physics, which involves changing relations between bodies within a frame. However, the fact that music is
perceived as mobile, that music ‘moves’ us, and that in classical music the parts of a piece are called
‘movements’, suggests that movement in physical and in musical ‘space’ can inform each other (think of
how lm music relates to movement on the screen, or dance music relates to dance). The notion of speed is,
p. 28 for example, common to both kinds of movement, and has to do with rates of change that are common to
how time is divided in both spatial and musical frames.

In Descartes’s example, the temporal occurrence of the elements of the song has to be converted into a C2.P9
connected whole—which is in some respects akin to the space of a painting—if the song is to be grasped as
a whole in the way required, for example, by a successful musical interpretation. The way in which we use
ideas from space, like movement, to grasp forms of time has to do with the need for identities and
similarities in relations between di erent elements of a series of temporally occurring phenomena. Because
they occur at di ering moments in time, the acoustic elements of music cease to exist as perceptible
phenomena once they have sounded, and so have to take on an idealized form of existence in memory for us
if they are to be connected. Rousseau suggests how music and time can relate in apparently paradoxical
ways:

It is one of the great advantages of the musician to be able to paint things that one could not hear, C2.P10
while it is impossible for the painter to represent things one could not see; and the greatest wonder
of an art which only acts through movement is to be able to convey by movement the very image of
repose. Sleep, the calm of the night, solitude, and even silence enter into the tableaux of music.

(Rousseau n.d.: Kindle Locations 27389–27391)

For this to be the case, the sense of time in music cannot be determined by music’s chronological movement C2.P11
from beginning to end; so the question is how musical movement comes to signify something like its
opposite. Music is expressed through a kind of temporal motion, and motion can involve contrasts. Intense
expression of movement conveyed by rapidity of changes in the music can mean that less intense motion
will be experienced as stillness. Analogously, the experience of silence is not the experience of nothing, but
of an absence of sound that signi es by its contrast with the presence of sound, and with the nature of the
preceding and succeeding sound. In this sense, the silences in music can be as signi cant as the sounds: the
‘pause’ after the explosion of orchestral intensity in the second movement of Schubert’s ‘Great’ Symphony
No. 9 in C major, D 944 (c.1825–8) has its own unique signi cance, as does the pause in Beethoven’s Egmont
Overture Op. 84 (1809–10), even though they just consist of ‘empty’ time. If there were only empty time,
there would, in one sense at least, be no time at all: time requires relations to be intelligible.
Rhythm, Self-Consciousness, and Music C2.S4

The presence and absence of sounds across time can just be the occurrence of random physical phenomena, C2.P12
so the key issue is how such phenomena become signi cant by being linked together into something
understood as meaningful. This situation applies both to music and to language: for something to be a note
p. 29 or a word, it must relate to other things in a way which generates sense. For this linking to be possible,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
there has to be something for which di erent sonic events are connected, and this something has itself to be
continuous between the events, even as it experiences the di erence of the sounds. The combination of
di erent phenomena which become intelligible to something which must itself remain identical is,
importantly, a way of describing self-consciousness. Kant talks of what he terms the ‘transcendental unity
of apperception’, an ‘ “I think”, that must be able to accompany all my representations’ (Kant 1968: 131),
which cannot be ‘as di erently multi-coloured as I have ideas that I am conscious of’ (p. 134) without
experience disintegrating and becoming unintelligible. For the world to be intelligible ‘I am conscious of
myself in general in the synthesis of the multiplicity of representations, therefore in the synthetic original
unity of apperception’ (p. 157). This is for Kant ‘the highest point to which one must attach all use of the
understanding, even the whole of logic’ (p. 134).

The initial point of this with respect to music and time is that, in Kant, even for natural scienti c accounts C2.P13
of chronological time to be possible, the relationship between the identity of the subject and the di erence
in what it apprehends must be in play (which means he sees time as based on our ‘internal sense’). Kant is
concerned, in the discussion of transcendental unity, with what makes knowledge possible at all. Knowledge
depends on active synthesis in judgements, on ordering the relationships between distinct appearances by
subsuming them under categories and concepts. This necessarily involves time, because a causal relation,
for example, entails the before and after of the appearances which are the data for the relation. The
transcendental unity of the subject is therefore required as the basis of cognitive synthesis. The question is
then why music takes on the kind of—not exclusively cognitive—meaning it does for the subject, by being
experienced as a connected series of temporal phenomena.

Identity in di erence is one way of characterizing rhythm: the di erent beats of a rhythmic pattern must be C2.P14
apprehended as being in one sense the same for rhythm to be such, rather than just random noises. This
alone, however, does not explain how rhythm becomes something meaningful. It is notable that rhythm
becomes more of an object of philosophical attention from the second half of the eighteenth century
onwards. This happens just as music moves from being a subordinate form of art to being regarded by many
as the highest form of art (see Bowie 2007; Dahlhaus 1974), and as a more intense awareness of the role of
time and history in how the world is understood also emerges. In his 1802–3 Philosophy of Art, F. W. J.
Schelling claims that rhythm is the ‘music in music’: rhythm is the ‘imprinting of unity into multiplicity’
(Schelling 1856–61: 492), which means that harmony and melody can also be conceived of as involving
rhythm. The idea of the unity of a multiplicity of elements is central to the aesthetics of classicism from the
end of the seventeenth century (see e.g. Baeumler 1967), but such aesthetics rarely takes seriously the
temporal dimension of uni cation of di erence.

The new attention to temporality in this period relates both to the incursion of secular time (which is in part C2.P15
an e ect of the spread of capitalism with its demands to regulate work time) in contrast to Christian and
Platonic conceptions, and to developments in mathematical physics after Newton. The higher status
p. 30 generally attributed to music, and the rapid changes in music from the second half of the eighteenth
century onwards, also suggest an awareness of a change in the awareness of time. Schelling points to this
change when he characterizes rhythm as essential to meaning, seeing it as ‘the transformation of a
succession which is in itself meaningless into a signi cant one’ (Schelling 1856–61: 493). The idea of the
initially meaningless nature of the succession of phenomena is decisive, because it indicates the need to
establish meaning in phenomena which may otherwise seem devoid of it once theological guarantees are
put in question. This takes us back to Kant’s focus on the subject’s role in establishing intelligibility, and the
attendant sense that meaning, including the meaning of time, cannot be assumed to pre-exist our
interactions with the world. An important aspect here is the di erence between knowledge of objects which
arises through the cognitive activity of the subject, where the truth of the object is presumed in some sense
to be already ‘there’ before it becomes known—the language of scienti c claims about nature being
tenseless—and forms of meaning which cannot be said to be ‘there’ before the activity of the subject in the
world, because they add to what the world can be understood to be.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
ʻSchematismʼ and Music C2.S5

One of the most di cult issues in Kant’s project of describing what makes knowledge possible is how to C2.P16
connect ‘pure’ forms of thought—i.e. forms not based on empirical data, which establish the identity of
things across time—with the endless diversity of what we experience in perceiving the world. The former
cannot themselves be endlessly diverse, because we could not bring any order to the world if they were; but
if the latter is contingent and irreducibly particular, it would seem impossible to bring it under forms of
identity. As we just saw, rhythm relies on ‘imprinting of unity into multiplicity’, and this leads to a key
point. Kant introduces the notion of ‘schema’ to bridge the gap between the ‘spontaneous’ operations of
pure thought and the ‘receptive’ apprehension of ‘empirical multiplicity’. For empirical cases of perception
of what Charles Taylor terms ‘independent objects’ a schema is what allows us, for example, to see both a
chihuahua and a Great Dane as dogs, despite all the massive di erences in how they look. The schema, then,
bridges the gap between the general concept and the particulars that come under it. F. D. E. Schleiermacher
suggests that a schema is a ‘shiftable image’ (in Frank 1989: 28), that can accommodate widely varying
instances of related things and unify them under a concept.

For Schelling ‘The schema […] is not an idea [Vorstellung] that is determined on all sides, but an intuition of C2.P17
the rule according to which a particular object can be produced’ (Schelling I/3: 508). He does not say it is
‘knowledge of a rule’, because that would be a concept, whereas what is in play is what makes the
application of concepts possible in the rst place, and he will link this to music and time. If it were a rule,
Kant’s point about judgement would apply—there would need to be a rule for the application of the rule,
p. 31 and so on, which leads to a regress that makes intelligibility impossible. Kant’s famous remark that
schematism is ‘an art hidden in the depths of the human soul’ makes it clear that what is in question here
are kinds of intelligibility that cannot themselves be further explained, because without them there can be
no explanations. David Bell argues in relation to schematism: ‘That our thought conform[s] to the rules,
principles, concepts and criteria constitutive of objectivity, but that it also be grounded in a spontaneous,
blind subjective awareness of intrinsic but inarticulable meaning—these are not con icting requirements’
(Bell 1987: 241). It is perhaps unsurprising, then, that music becomes so signi cant in this period, because
its ‘meaning’ as something which can disclose aspects of the world is undeniable, even if its signi cance
resides not least in the fact that it cannot be fully cashed out into semantic terms of the kind used to
describe independent objects.

The link of music to these issues becomes more apparent when Kant sees time itself in terms of ‘schemata’. C2.P18
The categories, like causality, necessity, etc., that make knowledge possible rely on prior forms of
intelligibility that are also bound up with the signi cance of rhythm and other aspects of music. The schema
of ‘cause’ ‘consists in the succession of the manifold to the extent to which it is subordinated to a rule’; of
‘reciprocity’ is ‘the simultaneity of the determinations of one substance with those of the other according to
a general rule’; of ‘reality’ is ‘existence of a thing at a certain time’; of ‘necessity’ is ‘the existence of an
object at all times’ (Kant 1968b: 184; 1968a: 145). Schemata are ‘therefore nothing but determinations of time
a priori according to rules’ (1968a: 145). These can be linked to music in the following ways. Succession is
made rule-bound, rather than merely arbitrary, by the category of cause, and something analogous applies
to the way rhythm makes random succession into a coherent sequence which can be felt as such, without it
needing to be conceptualized as such. Reciprocity can function as part of a description of harmony applied to
chords whose tones occur simultaneously. Reality and necessity can be seen as underlying the relations of
notes in a piece that depends on tonality, where the sounding of a note at a certain time makes sense
because the tonality is in play at all times and is necessary for comprehending what is heard as music. The
idea of ‘determinations of time a priori according to rules’ can itself be taken as a characterization of rhythm.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
Rhythm, Philosophy, and Music C2.S6

The intriguing question here concerns the status of rhythm in relation to forms of thought. In one sense, as C2.P19
we have seen, rhythm must depend on the unity of the subject that experiences the di erent moments of a
rhythm as intelligibly connected. This might seem to suggest the primacy of idealization in making nature
intelligible. On the other hand, rhythm qua linking of moments into something meaningful is part of
existing in the world in ways which precede abstract thinking. John Dewey suggests: ‘What is not so
generally perceived is that every uniformity and regularity of change in nature is a rhythm. The terms
p. 32 “natural law” and “natural rhythm” are synonymous’ (Dewey 1980: 149). The presence of rhythm in
nature allows one to ask about the genesis of schemata—an issue which Kant does not think belongs to
philosophy, but which will become central to subsequent thinkers like the German Idealists and Romantics,
and Heidegger.

If we did not inhabit a world which itself always already involves rhythm, such as the beating of one’s heart, C2.P20
one’s breathing, or the changes from day to night, birdsong, etc., what would make the world intelligible at
all in the rst place? Bell’s idea that objectivity is grounded in ‘spontaneous, blind subjective awareness of
intrinsic but inarticulable meaning’ makes sense in terms of rhythm qua transformation of meaningless
succession into something pre-theoretically meaningful. The use of rhythm in work-song to render
unpleasant physical work more tolerable suggests a kind of meaning that enables people to inhabit an
otherwise oppressive world. Such meaning is conveyed by making repetition into a means of escaping the
pressure of physical reality. At the same time, rhythm can itself become oppressive if it merely enforces
regularity. This creates the need for ways of countering rigidity in rhythm that play a major role in the
history of music, notably in the development of jazz, where the demand for ‘swing’ and the increasing
complexity of the division of rhythm are essential to keeping rhythm ‘alive’. Schelling makes the point very
clearly: ‘In everything which is in itself pure identity of activity man seeks […] driven by nature, to establish
multiplicity and variety through rhythm. We cannot tolerate uniformity for very long in everything that is in
itself without meaning, for example in counting, we make periods’ (Schelling 1856–61, I/3: 493).

The idea of rhythm as an essential aspect of how to make sense of the world is developed by the early C2.P21
German Romantic thinker Friedrich Schlegel, who goes so far as to make rhythm the condition of the rst
development of philosophy in Greece. For Schlegel, the initial state of human awareness involves
unarticulated feelings, of the kind which are shared by animals. Thought brings with it a sense of
limitlessness, and this needs to be checked if it is not to become inarticulate and generate disorientation.
Ecstatic states that he sees as characteristic of the early human responses to the world have to be controlled
by developing forms which structure them. He suggests that the people who generated the culture of this
period, from dance, to poetry and philosophy, were:

full of the living idea of an incomprehensible in nity. If this idea is the beginning and end of all C2.P22
philosophy; and if the rst inkling of it expresses itself in Bacchic dances and songs, in
inspirational customs and festivals, in allegorical images and poems; then orgies and mysteries
were the rst beginnings of Hellenic philosophy; and it was not a happy idea to begin philosophy’s
history with Thales, and to make it suddenly appear as if out of nothing.
(Schlegel 1988: 10)

Developing a culture which can channel a sense of the unbounded into meaningful forms involves C2.P23
overcoming the state generated by the feeling of limitlessness that is the condition both of the free
development of thought and of the anxiety to which this gives rise. This development depends on rhythm.

p. 33 Schlegel’s idea is in one sense speculative anthropology, but the role of rhythm in calming young infants C2.P24
and giving them the sense of a structured world points to something fundamental about rhythm’s role in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
how we come to inhabit a meaningful world where our thoughts and feelings make sense, because it
involves forms of order. Thought, Schlegel maintains, requires the ‘drive to hold fast a feeling [Emp ndung]
for oneself and to repeat it’ (Schlegel 1988: 13). This makes possible the beginning of the ‘poetic capacity of
humankind’: ‘for only by sensuous limitation and sensuous distribution of the material of communication,
by rhythm, which in the case of the wild man therefore does not belong to excess but to need, can feeling […]
be expanded into a lasting and more universal e ectiveness’ (p. 13). As such: ‘rhythm in this childhood of
the human race is the only means of xing thoughts and disseminating them’ (p. 16). In The Gay Science,
Nietzsche advances similar ideas in a discussion of the origins of poetry in rhythm: ‘With [rhythm] one
could do everything: magically encourage work; compel a god to appear, to be close, to listen; to order the
future according to one’s will; discharge one’s own soul of some excess (of fear, mania, sympathy, desire for
revenge)’ (Nietzsche n.d.: Kindle Locations 25182–25184). Dewey suggests the scope of such ideas when he
maintains that ‘a common interest in rhythm is still the tie which holds science and art in kinship’ (Dewey
1980: 150), both seeking to introduce order where it may appear to be lacking.

Rhythm is both somatic, giving rise to pleasure and displeasure, and, when it has entered the space of C2.P25
human culture, normative: it involves ways of getting it right and getting it wrong. Time is given pre-
conceptual shape and signi cance by rhythm, and rhythm can be re ned and di erentiated in novel ways.
The connection between time, self-consciousness, and rhythm is developed in a variety of ways by German
Idealist and Romantic philosophers. The core idea here is the idea of negation as being fundamental to how
the world makes sense. Time is seen as a process of negation, in which each moment negates the preceding
moment, making it possible for moments to be determined by their relations to what they are not, and
giving rise to the issue of how to retain what has happened, in order to overcome the disappearance which is
the appearance of time. Schelling maintains that ‘time is itself nothing but the totality appearing in opposition
to the particular life of things’ (Schelling 1856–61, I/6: 220): the particular is always nite, but we would not
be aware of this without the sense of what goes beyond the particular that enables nite moments to be
related.

Time, Negation, and Music C2.S7

Something analogous to this is enacted in music, where the particular moments of the music only make C2.P26
sense as part of a greater whole. In some late eighteenth- and nineteenth-century symphonic movements,
like Beethoven’s, the concern is to integrate more and more diverse musical material into a dynamic
totality. This involves a constant forward movement which is structured by the forward momentum being
p. 34 manifested in repeated thematic material that is built from a small number of patterns of notes. The idea
of temporal di erentiation built on repeated patterns here testi es both to a sense of temporality as
inexorable movement and to a contrary concern to structure the movement of time in a meaningful way in
order to transcend time. Adorno (1993; see also Bowie 2013) points to analogies between such music and
Hegel’s philosophy. In Hegelian terms time is a structure of negations, in which the past is not the present
or the future, the present not the past or the future, and the future not the present or the past. This ideal
structure, the totality of these relations that is required to think about time, is itself non-temporal, and
without it time would make no sense.
The ideal structure of time takes us back to what Kant saw in terms of the transcendental unity of self- C2.P27
consciousness. Schelling connects this to music, echoing his remarks about rhythm: ‘the principle of time in
the subject is self-consciousness, which is precisely the institution [Einbildung] of the unity of
consciousness into multiplicity […] From this the close relationship of hearing in general, and of music and
speech in particular, to self-consciousness can be understood’ (Schelling 1856–61, I/5: 491), and he links
language and schematism: ‘From this necessity of schematism we can infer that the whole mechanism of
language must rest upon schematism’ (p. 509). With respect to music and self-consciousness, Hegel

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
discusses how the musical note (the same is true of the spoken word) is produced by the vibration of a
material body which occupies a part of space, vacates it, and returns to it. The oscillation of the body
producing the note corresponds to ‘this negative sensuousness which, as it is, is not, and in its not-being
already produces its being again’; ‘the restless self-negation’ which is the vibration is the ‘emergence’
(Hegel 2003: 43) of what is negated (i.e. the static body) as something intelligible. The implication is that
without such movement, there would be no intelligible time at all.

Hegel claims that music demands the division of time because ‘time stands in the closest relationship with C2.P28
the simple self which hears and should hear its inner-self in the notes’ (Hegel 1965: 283). Time, as Schelling
also noted, can just be ‘empty progression’ (p. 284), and the self must overcome this because its essential
nature is ‘return to self’ (p. 283): this constitutes the identity without which time would be ‘empty
progression’. By returning to itself, the subject ‘interrupts the determination-less sequence of temporal
points, makes incisions in the abstract continuity’, so that the I ‘remembers itself and nds itself again’ (p.
284) in music. The temporal order of music is not given in nature, and the satisfaction music gives by
making sense of time belongs ‘neither to time nor to the notes as such but is something that only belongs to
the I’ (p. 285). The I establishes a sense of its own meaning, its ‘feeling of self’ (p. 283), via the engagement
with objective sounds that become more than empty progression.

What occurs both on the side of the subject and of the object (in this case the vibrating body) is a movement C2.P29
of negation and overcoming of that negation that is fundamental to the world’s intelligibility. At issue here
is not the world as a totality of objects, but the world as a dynamic context of meaning, where movement is
fundamental both to things and to the subject that engages with them. This is well summed up in theologian
p. 35 and philosopher F. D. E. Schleiermacher’s remark in his unjustly neglected Aesthetics: ‘the physiological
basis of rhythm […lies] in the movements of life themselves. The connection of artistic productivity with the
movements of self-consciousness which are so directly linked to the activity in the movements of life is,
accordingly, unmistakably the main factor in musical production’ (Schleiermacher 1842: 393). Such
assertions need not result in substantive metaphysical claims about what time and music are, qua objects of
philosophy or science, but rather orient themselves in terms of what people do with music and how music
can a ect us as temporal beings. In this respect, Schleiermacher, as he so often does, adumbrates ideas
which will only come to be fully developed much later.
Time and Music as ʻDimensions of Our Beingʼ C2.S8

Merleau-Ponty argues, ‘Time is not […] a real process, an e ective succession which I would limit myself to C2.P30
registering. It is born from my relation to things’ (Merleau-Ponty 1945: 471), so ‘time is not an object of
knowledge, but a dimension of our being’ (p. 475). The relations between time and music suggest new
directions for philosophy which question whether its predominant concern should be with knowledge of the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
objective world. The phenomenological approach initiated by Edmund Husserl and Martin Heidegger, and
developed by Merleau-Ponty, sees the core task rather as the understanding of meaning and ‘sense’ (which
in both the French sens and the German Sinn has the connotation of ‘direction’) by attending to how we
inhabit the world as embodied, active beings. In the word ‘sense’ ‘we nd the same fundamental notion of a
being oriented or polarized towards what it is not, and so we are always led towards a conception of the
subject as ek-stase [which has the connotation of standing outside/beyond itself] and a relation of active
transcendence between subject and world’ (Merleau-Ponty 1945: 491). ‘Transcendence’ here denotes the
way the subject moves beyond itself into the world in order to make sense possible. The forms of that sense
include cognitive results that describe an objective world in causal terms, but these depend on prior forms of
sense that are always already in play because of the temporalized nature of human existence.

Time as ‘a dimension of our being’ is manifest in the forms that articulate time, which endow it with C2.P31
meaning. The fact that music takes on a more central role in the modern period relates to how religious
contents are modi ed and replaced by new ways of seeking to locate people in a meaningful world. This is
not least because music itself is also ‘a dimension of our being’, as what we saw in relation to rhythm
suggests. The musicologist Heinrich Besseler, who studied with Heidegger, terms music a ‘Weise
menschlichen Daseins’, a ‘manner of human existence’, where the word ‘Weise’ also has the older
connotation of ‘song’ or ‘melody’, thus making our very being ‘melodic’ (Besseler 1978: 45). Henri Bergson
p. 36 sees time as a dimension of our being in terms of what he calls ‘duration’: ‘a continuation of what is no
more in what is’ (Bergson n.d.: Kindle Location 17966). He illustrates this with the analogy of hearing a
melody, which would cease to be a melody if it were decomposed into its parts:

our internal duration, viewed from the rst to the last moment of our life, is something like this C2.P32
melody. Our attention can turn away from it and consequently from its indivisibility; but, when we
try to cut it, it is as if we passed a blade abruptly through a ame: we only divide the space occupied
by it.

(Bergson n.d.: Kindle Locations 17976–17979)

Our insight into the nature of things present in duration depends on ‘intuition’: ‘The insights achieved by C2.P33
intuition resist linguistic expression, linguistic expression presupposing as it does the abstractions of
analysis. So metaphysical insights must likewise resist linguistic expression on this conception’ (Moore
2012: Kindle Locations 12812–12815). The use of music in such contexts seems not to be fortuitous: if time
cannot be fully grasped by analysis, forms of articulation which are inherently temporal, make sense, and
cannot be fully comprehended in concepts, may reveal things that are not accessible to discursive forms.

Rather than being regarded just as a physical constant or a relation between things (see Rovelli 2018, who C2.P34
shows that there is no consensus on what time is in contemporary physics), time itself can also be seen
historically, in terms, for example, of how the forms in which it is manifested may be either liberating or
oppressive. In notes relating to his work in the early 1870s on Wagner and Greek tragedy, Nietzsche talks of
how the Takt, which means ‘beat’, ‘cadence’, ‘bar’, ‘is to be understood as something fundamental—i.e. the
most primary form of the sense of time [Zeitemp ndung], the form of time itself’ (Nietzsche n.d.: Kindle
Locations 55732–55733). In a letter to Carl Fuchs in 1888 Nietzsche comments, suggesting how the
meanings of forms of temporality change, that:
Our [modern] rhythm is a means of expression of a ect: ancient rhythm, time-rhythm, has, in C2.P35
contrast, the task of dominating a ect and eliminating it to a certain degree […] In the ancient
understanding rhythm is morally and aesthetically the reins which are put on passion. In short:
our kind of rhythm belongs to pathology, the ancient to ‘ethos’.

(Nietzsche n.d.: Kindle Location 176025)

Such di ering perspectives on rhythm raise the question as to whether, underlying the di erences, there C2.P36

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
really is time as a unitary phenomenon that is just articulated in di erent ways. Merleau-Ponty suggests
why the unitary notion of time is questionable:

Everyone talks about time [parle du temps] in the singular and not in the way the zoologist talks of C2.P37
the dog or the horse, in the sense of a collective noun, but in the sense of a proper name. […]
Everyone thinks there is a single concrete entity there, present in all its manifestations like a
person is present in each of their words. One says there is one time in the way one says that there is
p. 37 a jet of water; the water changes and the jet of water remains because the form is preserved; the
form is preserved because each successive wave takes up the functions of the preceding wave.

(Merleau-Ponty 1945: 482)

The jet of water echoes the idea of rhythm, in which di erent moments are seen as the same once they are C2.P38
apprehended as having a form. The ‘primordial experience’ we have of time ‘is not for us a system of
objective positions across which we pass, but a moving milieu which distances itself from us, like the
landscape in the window of the carriage’ (Merleau-Ponty 1945: 480). Establishing temporal ‘content’, as
opposed to merely abstract time as an accumulation of identical units, is only possible because of the (in
some sense timeless) identity of a subject which sustains the past in existence through memory and
anticipates the future: ‘You have to understand time as a subject and the subject as time’ (Merleau-Ponty
1945: 483). Contemporary physics suggests as well that time cannot be seen as unitary, and once again
music is used to understand whatever it is that time is: ‘Time has lost its rst aspect or layer: its unity. It has
a di erent rhythm in every di erent place and passes here di erently from there. The things of this world
interweave dances made to di erent rhythms’ (Rovelli 2018: Kindle Locations 181–183). The disintegration
of the idea of a uni ed time that is proposed by some contemporary physicists brings them closer to certain
key philosophical ideas in the European tradition.

Heidegger’s radicalization of philosophical approaches to time sees time itself as the condition of possibility C2.P39
of an intelligible world, and as essential to the meaning of being. The essence of our being is its nitude, and
our understanding of being must be based on how that nitude underlies the ways in which sense is made of
the world, rather than aiming to arrive at a ‘Platonic’ timeless account of that essence that takes us beyond
nitude. As Kant maintains, whereas a divine intelligence would be timeless, seeing the truth of what we
experience temporally all at once in a ‘total present’, and would itself create the world’s intelligibility in
creating the world, we can only arrive at knowledge of the world in the temporal form in which the world
manifests itself and which we order in cognitive and aesthetic ways. We saw how the conditions of
knowledge could be said to have their origin in the prior, non-conceptual forms of intelligibility essential to
music, and Heidegger makes Kant’s ‘schematism’ chapter central to his understanding of being. Because
the transcendental imagination is the condition of time taking intelligible form through apprehension,
reproduction, and recognition (which correspond to present, past, and future), it is ‘original time’
(Heidegger 1973: 201). In this sense—though Heidegger himself does not explicitly see it this way—time
can be understood, not as something whose sense is imitated by music, but as something whose sense is
inseparable from music.
Importantly, given the focus of so much modern philosophy on language, the same temporalizing aspect C2.P40
applies to verbal language, as Schelling suggested when he linked schematism and language. Martha
Nussbaum notes: ‘Musical works are somehow able—and, after all, this “somehow” is no more and no less
mysterious than the comparable symbolic ability of language—to embody the idea of our urgent need for
p. 38 and attachment to things outside ourselves that we do not control, in a tremendous variety of forms’
(Nussbaum 2001: 272). The fact that the possibility of language is itself predicated on the same fundamental
processes of schematizing past, present, and future, in order for signs to be iterable and signi cant in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
di erent contexts, again suggests how the prior, non-semantic kind of sense involved in music is
inseparable from any understanding of time.

Observation and Participation C2.S9

It is a commonplace of the philosophy of music—in the sense of philosophy which takes music as an object C2.P41
to be explained—that, as Peter Kivy puts it, ‘Music, of all the arts, is the most philosophically unexplored
and most philosophically misunderstood where it has been explored at all’ (Kivy 1997: 139). Something
related could be asserted of views of time and philosophy. In both cases the philosophical attempt to come
up with a theory of the object music, or the object time, seems to lead to endless disagreements, albeit
disagreements which can also produce valuable insights or clari cations of the problems. Such theorizing
relies on an observational stance towards music and time, and this stance is precisely what is questioned by
phenomenological approaches, like that of Merleau-Ponty, which relate to the traditions deriving from
Kant and German Idealism.

The important di erence is that phenomenological approaches think also in terms of a participatory stance. C2.P42
This assumes that what takes place in participation exceeds what can be objecti ed in theoretical accounts
of participation. Musicians gain the sense from what they do by inhabiting a practice that articulates time in
ways which can enrich the rest of our lives. They may well seek to theorize the practice, but do so on the
basis of a prior involvement that rst makes it worth attempting a theoretical articulation of the practice.
Pre-conceptual, participatory understanding of time tends to be neglected in accounts that focus on, say,
the perception of time in music. The point of the phenomenological approach is to bring to light things
which we always already do, but which can be obscured by objectifying forms of thought. We anticipate
actions on the basis of remembered experience, coordinate bodily actions, feel and respond to how our
moods change, etc. Aspects of all these forms of understanding are essential to music, and music can
enhance our ability to develop such forms. In these respects, time is not mysterious to us, just as it is not
mysterious to musicians or listeners to music when they are absorbed by music’s articulation of time. The
genesis of forms of sense at issue here is a result of our interactions with a world which, as we saw, can itself
be seen as rhythmically structured.

Such re ections on music and time should not exclude the observational perspective and its attendant C2.P43
analytical modes of theorizing. However, they do imply that without what is inherent in participation,
questions about time and meaning of the kind essential to thinking about music risk losing their point.
p. 39 Moreover, the participatory approach opens a perspective in which understanding time is inseparable
from active involvement with music, which cannot just be replaced by theoretical re ection. By adding to
our possibilities of articulation and expression, music can add to the sense we can make of time in a manner
which is thoroughly philosophical, if we conceive of philosophy in terms of making wider sense of the world
we inhabit. This is especially the case if we think of the ‘philosophy of music’ as the philosophy which
emerges from music, rather than making music its object. Why otherwise would the drive for innovation in
rhythm and melody which is part of the development of jazz have such signi cant cultural e ects? Why is it,
as Proust’s À la recherche du temps perdu (1913–27) shows, sometimes only music that can give access to past
memories in a way which recalls their speci c intensity? Why, for many secular people, can engagement
with music play a role that shares certain features with participation in religious ritual with its speci c
forms of temporality? Rather than assuming that philosophy should aim to achieve a comprehensive
explanatory account of music and time, it may sometimes make more sense for philosophy to point to what
music reveals about time which can only be grasped by active participation in music.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
References C2.S10

Adorno, T. W. (1993). Beethoven: Philosophie der Musik. Suhrkamp. C2.P44


Google Scholar Google Preview WorldCat COPAC

Auerbach, E. (1953). Mimesis: The representation of reality in Western literature. Princeton University Press. C2.P45
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
Baeumler, A. (1967). Das Irrationalitätsproblem in der Ästhetik und Logik des 18. Jahrhunderts bis zur Kritik der Urteilskra . C2.P46
Wissenscha liche Buchgesellscha .

Besseler, H. (1978). Aufsätze zur Musikästhetik und Musikgeschichte. Reclam. C2.P47


Google Scholar Google Preview WorldCat COPAC

Bell, D. (1987). The art of judgement. Mind 96(382): 221–244. C2.P48


Google Scholar WorldCat

Bergson, H. (n.d.) Oeuvres complètes (Annoté et illustré). Kindle Edition. C2.P49


Google Scholar Google Preview WorldCat COPAC

Bowie, A. (2007). Music, philosophy, and modernity. Cambridge University Press. C2.P50
Google Scholar Google Preview WorldCat COPAC

Bowie, A. (2013). Adorno and the ends of philosophy. Polity. C2.P51


Google Scholar Google Preview WorldCat COPAC

Dahlhaus, C. (1974). Zwischen Romantik und Moderne: Vier Studien zur Musikgeschichte des späteren 19. Jahrhunderts. C2.P52
Musikverlag Katzbichler.

Dahlhaus, C. (1978). Die Idee der absoluten Musik. Deutscher Taschenbuch-Verlag. C2.P53
Google Scholar Google Preview WorldCat COPAC

Descartes, R. (1987). Abrégé de musique: Compendium musicae. PUF. C2.P54


Google Scholar Google Preview WorldCat COPAC

Dewey, J. (1980). Art as experience. Perigee. C2.P55


Google Scholar Google Preview WorldCat COPAC

Frank, M. (1989). Das Sagbare und das Unsagbare. Suhrkamp. C2.P56


Google Scholar Google Preview WorldCat COPAC

Hegel, G. W. F. (1965). Ästhetik (vols 1 and 2, ed. F. Bassenge). Aufbau. C2.P57


Google Scholar Google Preview WorldCat COPAC

Hegel, G. W. F. (2003). Vorlesungen über die Philosophie der Kunst. Meiner. C2.P58
Google Scholar Google Preview WorldCat COPAC

Heidegger, M. (1973). Kant und das Problem der Metaphysik. Klostermann. C2.P59
Google Scholar Google Preview WorldCat COPAC

Kant, I. (1968a). Kritik der reinen Vernun , Werkausgabe 3. Suhrkamp. C2.P60


Google Scholar Google Preview WorldCat COPAC

Kant, I. (1968b). Kritik der reinen Vernun , Werkausgabe 4. Suhrkamp. C2.P61


Google Scholar Google Preview WorldCat COPAC

Kivy, P. (1997). Philosophies of arts: An essay in di erences. Cambridge University Press. C2.P62
Google Scholar Google Preview WorldCat COPAC

Merleau-Ponty, M. (1945). Phénoménologie de la perception. Gallimard. C2.P63


Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468911 by National Science & Technology Library user on 26 May 2023
Moore, A. W. (2012). The evolution of modern metaphysics: Making sense of things. Cambridge University Press. Kindle Edition. C2.P64
Google Scholar Google Preview WorldCat COPAC

p. 40 Nietzsche, F. (n.d.). Sämtliche Werke. Oregan Publishing. Kindle Edition. C2.P65


Google Scholar Google Preview WorldCat COPAC

Nussbaum, M. (2001). Upheavals of thought: The intelligence of emotions. Cambridge University Press. C2.P66
Google Scholar Google Preview WorldCat COPAC

Rousseau, J.-J. (n.d.). Oeuvres complètes, nouvelle édition enrichie. Arvensa. Kindle Edition. C2.P67

Rovelli, C. (2018). The order of time. Penguin Books. Kindle Edition. C2.P68
Google Scholar Google Preview WorldCat COPAC

Schelling, F. W. J. (1856–61). Sämmtliche Werke, ed. K. F. A. Schelling. Cotta. C2.P69


Google Scholar Google Preview WorldCat COPAC

Schlegel, F. (1988). Kritische Schri en und Fragmente 1–6. Schöningh. C2.P70


Google Scholar Google Preview WorldCat COPAC

Schleiermacher, F. D. E. (1842). Vorlesungen über die Ästhetik. Reimer. C2.P71


Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
CHAPTER

3 Forms of Time in Nineteenth-Century


C3 Music: Geology, the
Railway, and the Novel 
Lawrence Kramer

https://doi.org/10.1093/oxfordhb/9780190947279.013.7 Pages 41–56


Published: 08 December 2021

Abstract
European art music in the nineteenth century was characterized by both an expansion and a
contraction of the timescale typical of earlier periods. On the one hand there was an outpouring of
miniatures, primarily for piano; on the other there was a proliferation of instrumental works,
especially symphonies, lasting anywhere from 40 minutes to over an hour. Although it is possible to
refer these changes to developments in compositional technique, their wider signi cance derives from
the era’s production of several new and epoch-making forms of time––that is, of ways to conceive,
order, and experience time. Time literally changed during the nineteenth century, and music changed
along with it. The long span of geological ‘deep time’, the compressed and precisely measured time of
railway travel, and the temporal complexity of the multiply plotted novel all have musical parallels.
Robert Schumann’s Symphony No. 4 in D minor Op. 120 (1841), César Franck’s Symphony in D minor
(1888), and Frédéric Chopin’s Prelude No. 18 in F minor Op. 28 (1835–9) provide pertinent examples.

Keywords: time, temporality, miniature, expansion, deep time


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

Changing Times C3.S1

DURING the nineteenth century, art music got both longer and shorter. On the one hand, there was an C3.P1
outpouring of miniatures and stand-alone character pieces, primarily for solo piano. On the other hand,
there were symphonies running—expected to run—from 40 minutes to an hour and more. Why?

There are certain standard musical explanations based on changes in compositional technique: increasing C3.P2
reliance on the recycling of short motives, the rise of developing variation, thematic transformation and
apotheosis, the melodic and harmonic expansion of sonata form. But these devices are less the causes of the
era’s new temporalities than the means by which those temporalities were echoed and reshaped. The more
informative answer (though it does not, as none can, quite rise to the level of causation) comes from
temporal changes that rippled across the whole cultural–material spectrum. These were changes that linked
the time of music to the time of the lifeworlds of those who composed and played and listened. Time and
music changed together and changed each other.

Music has no xed temporal form. It can realize virtually any mode of temporality you like; it is a kind of C3.P3
existential timepiece. During the nineteenth century, at least three new or transformed modes of time

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
together with their broad cultural impact sound out clearly in music: deep or epochal time, the time of
cosmology and geology, fossils, and evolution; novelistic time, with its multiple plots, complex histories,
p. 42 haunting secrets, and layers of contingency; and railway time, especially the lived time of the railway
journey. These forms of time correlate, respectively, with symphonic expansiveness, doubling or more than
doubling the span of most earlier symphonies; with the large-scale multi-movement designs that
increasingly frame and ll the symphonic expanses; and, at the other end of the spectrum, with the
proliferation of miniatures, some of them lasting barely a minute, primarily in collections for piano.

The expanding timescale has obvious precedents in Beethoven, but they are misleading because their C3.P4
temporal modalities are both unrepresentative and too representative. Most of Beethoven’s works on the
expansive scale (the Piano Trio in B♭ major Op. 97 ‘Archduke’ (1811) and Piano Concerto No. 5 in E♭ major
Op. 73 ‘Emperor’ (1809–11) among many others) did not become models for his successors, except insofar
as they helped legitimate the scale itself. The role of model fell primarily to two very famous symphonies:
No. 5 in C minor Op. 67 (1804–8) for its explicit continuity across multiple movements, and No. 9 in D minor
Op. 125 (1822–4) for the explicit narrativity introduced when the nale begins with a review of music from
the preceding movements. But these models were received less as paradigms than as problems. They could
not be emulated unless they were reinvented. The questions of what music should do in expanded time and
how it should do it remained open.

Beethoven also provided precedents on the contracted scale, but the nineteenth century largely ignored C3.P5
them. Examples include the minimalist scherzo of the Violin Sonata No. 5 in F major Op. 24 ‘Spring’ and the
miniature movements in the trio of late string quartets with more than four movements—No. 13 in B♭ major
Op. 130 (1826), No. 14 in C♯ minor Op. 131 (1826), and No. 15 in A minor Op. 132 (1825). Here, too, there is a
kind of false start. The quartets in particular are works on the expanded scale. Their miniature movements
contract as a concomitant of the expansion that envelops them. They thus imply a mode of time quite unlike
that of the century’s paradigmatic miniatures, Chopin’s Preludes Op. 28 (1835–9). Beethoven’s tiny quartet
movements identify their brevity with compressed totality rather than with fragmentation. Their lyric time
is not self-su cient. Instead, it tends to form a preface or passageway to musical narrative, something that
happens in all three of the quartets but most emphatically in Op. 131, which ends with what should have
been its rst movement.

Wagner is another matter. He is certainly the century’s most in uential innovator and practitioner of the C3.P6
expanded timescale, regardless of the di erence in kind between opera and symphony. Wagnerian time is
not a product of duration measured by the clock; Meyerbeer’s operas, and others, are just as long. The key
factor is the unbroken chain of events celebrated or derided at the time by the term ‘endless melody’.
Primarily through Wagner, opera as concatenation becomes opera as continuity. Composers of symphonies
took note. But the expansive impulse also predates Wagner, with stirrings in the 1820s and 1830s in the
work of Schubert and Berlioz. The driving force here is not individual precedent but cultural persuasion.

Among canonical composers, Chopin, Mendelssohn, and Schumann were probably the most absorbed with C3.P7
the contracted scale (at least until one gets to the not-really-canonical Anton Webern many decades later).
This observation comes with the interesting wrinkle that the arc of Schumann’s career is essentially a
p. 43 movement from contraction to expansion: from collections of short (sometimes very short) piano pieces
in 1830s and songs in 1840, to large-scale chamber and symphonic works thereafter. Chopin studiously
avoided anything of the kind. The Preludes represent an extreme, but he preferred ‘small forms’ throughout
his career. Mendelssohn did not, but his eight volumes of Songs without Words (1829–45) were widely
popular.

Nineteenth-century compositions on the contracted scale di er from the short pieces of the preceding C3.P8
century in two ways. First, they regularly seek critical distance from the genres to which they nominally
belong, if indeed they belong to any (they may well not). Second, they decline to stand as minor or lesser by

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
virtue of their short span. The rst di erence might be described by saying that the rule of genre could be
followed only if it was also broken. (In a literary context, Jacques Derrida called this principle ‘the law of
genre’: 1980) Chopin’s waltzes and nocturnes furnish prime examples. But the second di erence is perhaps
the more culturally resonant. The change is not, of course, the revelation that music could be short,
although there is a revelation that it could be very short and yet form a totality. The change, rather, is that
music’s claims to signi cance are not limited by brevity: small moments could have large signi cance.
William Wordsworth called such moments ‘spots of time’ and assigned them a power far out of proportion
to their swift, almost incidental, occurrence:

A virtue, by which pleasure is enhanced, C3.P9


That penetrates, enables us to mount C3.P10
When high, more high, and lifts us up when fallen. C3.P11

(The Prelude, XI: 265–268))

‘Such moments,’ he adds, ‘worthy of all gratitude, / Are scattered everywhere’ (pp. 273–274). Late in their C3.P12
lives, both Liszt and Brahms were drawn to the musical equivalent in concentrated piano works, many with
an elegiac cast (Liszt’s Nuages gris [Grey clouds] S.199 (1881), Bagatelle sans tonalité S.216a (1885), and
‘Abschied’ [Farewell] S.251 (1838–9), among others; Brahms’s series of piano pieces collected in Opp. 116–
19).

The Expanded Scale: Symphonic Novels and Symphonic Geology C3.S2

Novelistic time is distinctive to the symphony, from which it occasionally extends to large-scale concertos C3.P13
and chamber works. It does not, however, have anything to do with musicalized narrative, that is, with
programme music, the span of which is relatively short and the source of which is usually a poem, especially
a narrative poem. The symphony becomes novelistic in relation, not to topic, but to time. This relationship
will echo throughout the ensuing discussion.

Another point to observe is that theories of the novel in the nineteenth century rarely concerned themselves C3.P14
p. 44 with time. The aesthetic of the novel in the era of its ascendency focused primarily on the realistic
representation of character and social (dis)order. As Stendhal famously put it in a self-re exive passage of
The Red and the Black (1830), ‘A novel is a mirror walking along a highway’ (Stendhal 1989: 43). The driver of
novelistic length was the demand for comprehensiveness; each social echelon needed a plot of its own.

The recognition that this imperative results in a distinctive mode of time itself took time to develop. The C3.P15
musical variety of novelistic time absorbs a practice more than it hews to a principle. In some cases, it may
make just as much sense to say that the time of a novel is symphonic. One might even say that novelistic
time nds its rst theorization in the symphony. Nothing as simple as in uence or imitation is at stake
here. The issue is the cultural di usion of aesthetic predispositions, or what Jacques Rancière calls ‘the
distribution of the sensible’: the work of ‘the a priori system of forms that determines what presents itself to
sense experience’ (2006: 13). Rancière is primarily interested in the political stakes of this ‘delimitation of
spaces and times, of the visible and the invisible, of speech and noise’, but its cultural stakes are equally
fraught and equally sensory. Time in particular, as part of the nineteenth century’s distribution of the
sensible—in music and geology, the novel and the railway—becomes not only a primary medium of sense
experience but also one of its primary objects.

The association between the nineteenth-century novel and music on the expanded timescale does not begin C3.P16
with me; it begins with Robert Schumann. In his account of his 1838 discovery of the score of Schubert’s
Symphony No. 9 in C major, D 944 (1825–8) ‘buried in dust and darkness’ in the house of Schubert’s

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
brother Ferdinand, Schumann famously celebrates the ‘heavenly length’ of the hour-long symphony (the
phrase passed into legend). He compares the music to an imaginary novel by one of his own literary heroes:
‘[It is] like a hefty novel in four volumes, as if by Jean Paul, which also can never end, and for the best of
reasons, so as to let the reader keep on creating in its wake’ (wie ein dicker Roman in vier Bänden etwa von
Jean Paul, der auch niemals endigen kann, und aus den besten Gründen zwar, um auch den Leser hinterher
1
nachscha en zu lassen; Schumann 1854: 201).

This remark identi es what we would nowadays call immersiveness as a primary feature of work on the C3.P17
expanded scale. The listener, like the reader, continues the work of art—not the thing but the activity—in
becoming wholly absorbed in it. The nineteenth-century custom of publishing novels in instalments,
heightening expectation as one volume followed the next, becomes the template for listening to the four
extended movements.

Schumann subsequently continues the metaphor (though it is perhaps more than that), observing that the C3.P18
‘novel intricacies’ that abound in the music, extending throughout ‘the length and breadth of form’, may at
rst be disorienting, but that ‘a delightful feeling remains, as though we had been listening to a eeting tale
of fairies and enchantment. We feel that the composer has mastered his tale, and that, in time, its
connections will all become clear.’ The composer, present in his work in the manner of a self-re ective
omniscient narrator, weaves the strands of plot together on the basis of a principle that becomes clear only
in retrospect. He acts, in other words, almost exactly like the narrator of a multi-volume, multiply plotted
novel.

p. 45 Schumann’s descriptions of the symphony elsewhere in his article are consistently novelistic. He refers the C3.P19
music not to its customary province of ‘mere beautiful song, mere joy and sorrow, such as music has ever
expressed in a hundred ways’, but to the province of the long novel, ‘the outer world, sparkling today,
gloomy tomorrow, [which] often deeply stirs the feeling of the poet [Dichter] or musician’. The symphony
reveals itself as ‘sharp in detail’ and replete with ‘meaning everywhere’; it ‘lead[s] us into regions which, to
the best of our recollection, we had never before explored’. Schumann does not, however—and the point is
crucial—assign a narrative to the symphony or surmise a programme for it. He is not concerned with the
story it might tell but with the process of the telling. Above all he is concerned with the temporality of the
telling, the feel of time on the expanded scale. The alternation of mood from one day to the next, the
interplay of memory and discovery, the metaphor of a journey into an unknown region, all combine to
ground ‘the ctional character that pervades the whole symphony’ in the passage of time in long,
overlapping waves.

The symphony in which Schumann himself sought to realize this mode of time is, ironically, his shortest: C3.P20
No. 4 in D minor Op. 120, composed in 1841 and revised ten years later. The second version is my subject
here. The symphony lasts only a little over a half hour, but its timescale is nonetheless a long one. Although
there are four distinct movements, they are played without pause; the second and the third movements are
interlocked; the third and fourth are linked by a transitional passage based on the introduction to the rst;
and thematic material from the rst is recycled in the fourth. The symphony is thus more in one movement
than it is in four. It conforms to only one undivided ow of time: the long span. Schumann’s decision in 1851
to omit the pauses between movements that appear in the rst version, and to add the transition between
the third and fourth movements, counts as a recognition that the undivided ow had informed the music
from its inception.

The incessant cross-reference in Schumann’s Symphony No. 4 gives explicit form to the novelistic impetus C3.P21
that he had heard in Schubert’s Symphony No. 9. No listener is supposed to miss the looping and looming
that results. The music moves through time in the involuted manner of a multiply plotted novel, but without
making reference to any particular plots. The symphony reaches for narrative time, not narrative content.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
Nonetheless, to give its mode of time a distinctive feel, the work does one of the things that its era’s multi-
tiered novels typically do: it discloses a secret step by step.

The secret in this case is the value belonging to a sequence of three declamatory chords (at rst the same but C3.P22
subsequently varied, the real motif being the triple declamation). The chords themselves are clear; it is their
meaning that is at stake. Introduced in a subsidiary role about ve minutes into the rst movement, the
declamatory gure recurs in various forms throughout. In the scherzo it supports the wilfully crude
stamping of the main theme. In the transition to the nale it transforms itself into a series of ambivalent
brass fanfares—sacred or secular? solemn or stern?—while the introduction to the rst movement returns
with gathering intensity. The nale begins with a fortissimo statement of the gure for full orchestra. The
triple attack quickly emerges as the driving force of the movement and therefore of the whole symphony,
which ends after a headlong rush with one last fortissimo declamation. After long being hidden in the
p. 46 auditory equivalent of plain sight, the triple attack stands revealed as a primal burst of energy from
which the whole system of melodic cross-references can be imagined to proliferate across the long span.

It is important to emphasize that there is nothing esoteric about this. The process is transparent; it insists C3.P23
on being heard. The point is not to mystify time but to make its passage feel a certain way. The same
consideration will apply when we turn to Franck’s Symphony in D minor (1888) below.

Epochal time, or ‘deep time’ in John McPhee’s felicitous coinage (1982: 27), began in the late eighteenth C3.P24
century as a concept in geology, merged in the mid-nineteenth century with the time of Darwinian
evolution, and merged again, following the subsequent development of anthropology, with the concept of a
human prehistory preserved in the cultures of ‘primitive man’. The dissemination of these orientations was
international, but an emblematic turning point occurred in England with the publication of Charles Lyell’s
Principles of Geology in 1830–33. Lyell’s point of departure was the work of the Scottish geologist James
Hutton, who described a hitherto unthinkably vast timespan for the history of the earth, leaving the mind,
as the mathematician John Playfair famously observed (1830–33: 81), to ‘grow giddy with looking so far into
the abyss of time’. Lyell enlarged on the point (and its religious implications):

No small sensation was excited when Hutton seemed, with unhallowed hand, desirous to erase C3.P25
characters already regarded by many as sacred. ‘In the economy of the world,’ said the Scotch
geologist, ‘I can nd no traces of a beginning, no prospect of an end;’ and the declaration was the
more startling when coupled with the doctrine, that all past changes on the globe had been brought
about by the slow agency of existing causes. The imagination was rst fatigued and overpowered
by endeavoring to conceive the immensity of time required for the annihilation of whole
continents by so insensible a process […] Such views of the immensity of past time, like those
unfolded by the Newtonian philosophy in regard to space, were too vast to awaken ideas of
Sublimity unmixed with a painful sense of our incapacity to conceive a plan of such in nite extent.
Worlds are seen beyond worlds immeasurably distant from each other, and beyond them all
innumerable other systems are faintly traced on the con nes of the visible universe.

Apart from its eloquence, Lyell’s book is important because Darwin took it with him on the Beagle. For C3.P26
evolution to be conceivable, the epochal time required for it must be conceived rst.
The mapping of temporal expanses onto spatial vistas with which the passage from Lyell concludes became C3.P27
habitual as the century progressed. For many, however, the habit only sharpened the accompanying
anxiety, as the two literary examples to follow will illustrate. Baudelaire’s poem ‘The Abyss’ is one of the
rst to seize on the imagery of deep time, or rather to feel seized by it:

On high, down low, all around, the depth, the verge, C3.P28
p. 47 The silence, space hideous and captivating […] C3.P29

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
On the background of my nights God with a knowing nger C3.P30
Designs a multiform never-ceasing nightmare. C3.P31

(Baudelaire 1861 [2004])

Baudelaire’s nightmare translates the expanse of space into the abyss of time. Vertigo follows; every nite C3.P32
perspective becomes in nite in a never-ceasing series of repetitions: ‘I see only the in nite through all the
windows.’ The mind and senses reel with the same kind of incapacity that Lyell invokes to explain the
resistance to Hutton’s world picture.

Tennyson, in ‘In Memoriam’ (1850: no. 123), is even more explicit than Baudelaire about the way that the C3.P33
image of an external in nity recoils to become an internal gulf, a source of spiritual vertigo:

There rolls the deep where grew the tree. C3.P34


O earth, what changes hast thou seen! C3.P35
There where the long street roars, hath been C3.P36
The stillness of the central sea. C3.P37

The hills are shadows, and they ow C3.P38


From form to form, and nothing stands; C3.P39
They melt like mist, the solid lands, C3.P40
Like clouds they shape themselves and go. C3.P41

Tennyson’s stanzas collapse the immediate present and the geological past into a moving phantasmagoria C3.P42
that plays out, not in the external world, but in the mind and senses of the observer. The landscape appears
to change vertiginously but the changes occur only in the observer’s vertigo. Tennyson is mournful where
Baudelaire is appalled, but both poets register a version of the shock described by Lyell. For both, the
‘immensity of time’ leaves the imagination ‘fatigued and overpowered’. Tennyson, however, subtly allows
his verse to register the almost musical uidity of the vertigo induced by the contemplation of deep time,
and in that respect he hints at a di erent possibility of response.

Musically, the ‘dark backward and abysm of time’ (to recall Shakespeare’s The Tempest, the source of both C3.P43
Playfair’s description and Tennyson’s cloud imagery) seems to have registered less as a problem than as an
opportunity. Paradoxically, the expansion of time enabled a resonant form of modern awareness by
shrinking the magnitude of modern existence. In an era that felt increasingly encroached upon by its
geological, evolutionary, and anthropological vistas, the symphony found a means of transforming the
limitless span into a frame of reference. Music could make a vessel of the abyss.

Deep time, epochal time, primarily frames symphonic works in which the music is in continuous evolution, C3.P44
usually, though not always, across multiple movements. Most of these works date from the later years of the
long century, extending to the catastrophic break of 1914. Canonical instances spanning multiple
movements include Franck’s Symphony, Tchaikovsky’s Symphony No. 5 in E minor Op. 64 (1888), Mahler’s
p. 48 Symphony No. 2 (1888–94) and No. 5 (1901–2), and Sibelius’s Symphony No. 1 in E minor Op. 39 (1898–
9), No. 2 in D major Op. 43 (1901–2), and No. 5 in E♭ major Op. 82 (1915). Mahler’s Symphony No. 5 is an
especially revealing case because it drops melodic cross-reference after the second of its ve movements, as
if the way forward required creative forgetting rather than the recollection essential to the Second.

These works tend to amplify the end-weighting characteristic of the symphonic genre after mid-century. C3.P45
The nale of Vaughan Williams’s A Sea Symphony (1901–10) lasts for half an hour, long enough to equal or
exceed all four movements of any of Haydn’s London Symphonies. The work begins with two celebrated
gestures—a brass fanfare and an abrupt third-related key change—that reoccur across its 70-minute span.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
Like Mahler’s contemporaneous Symphony No. 8 in E♭ major (1906), A Sea Symphony uses voices
throughout. Whereas Mahler begins with the medieval hymn ‘Veni, Creator Spiritus’ and follows it with a
far longer setting of the nal scene of Goethe’s Faust, Vaughan Williams continuously sets poetry by Walt
Whitman in which the sea becomes a mirror of deep time. Whitman (2001: 218) replaces Baudelaire’s terror
and Tennyson’s mournfulness with rhapsodic a rmation and dissolves their tightly controlled lyric forms
into expansive free verse:

On the beach at night alone, C3.P46


As the old mother sways her to and fro singing her husky song, C3.P47
As I watch the bright stars shining, C3.P48
I think a thought of the clef of the universes and of the future. C3.P49
A vast similitude interlocks all, C3.P50
All distances of space however wide, C3.P51
All distances of time, C3.P52
All souls, all living bodies though they be ever so di erent, C3.P53
All nations, all identities that have existed or may exist, C3.P54
All lives and deaths, all of the past, present, future, C3.P55
This vast interlude spans them, and always has spanned, C3.P56
And shall forever span them and shall compactly hold and enclose them. C3.P57

Mahler’s even longer Symphony No. 3 (1893–6)—at roughly 100 minutes the longest symphony in the C3.P58
standard repertoire—frames four modestly scaled movements with two vast ones and stretches time still
further by making the second of these, the nale, an Adagio. (Mahler would do the same again in his
Symphony No. 9 (1908–9).) The fourth movement makes the symphony’s occupation of epochal time
explicit with a setting of a passage from Nietzsche’s Thus Spoke Zarathustra:

O man, take heed! C3.P59


What does the deep midnight say? C3.P60
‘I slept, I slept— C3.P61
from a deep dream I have wakened:— C3.P62
The world is deep, C3.P63
and deeper than day has thought. C3.P64
p. 49 Deep is its pain— C3.P65
Joy—deeper still than heartache. C3.P66
Pain says: pass away! C3.P67
But all joy wants eternity —. C3.P68
wants deep, deep eternity!’ C3.P69

(Nietzsche 1985)

For those wondering about the absence of Bruckner from this listing, su ce it to say that I hear his C3.P70
symphonies as aspiring to a static quasi-ecclesiastical time rather than to the dynamic epochal mode,
though the distinction is easier to draw than it is to maintain.
In all of these works, to ri on a famous line from Wagner’s Parsifal (WWV 111), time becomes spacious. For C3.P71
the most part, however—the Vaughan Williams and Mahler’s Symphony No. 3 are the notable exceptions—
these symphonies are not ‘about’ epochal time. Like Schubert’s Symphony No. 9 in Schumann’s conception
of it, which inhabits novelistic time without a narrative, these long-span works follow an epochal pulse
without re ecting on it. Deep time is not their topic but their horizon.

Franck’s Symphony is a particularly revealing instance; it inhabits both expanded timescales at once. The C3.P72

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
rst and longest movement is built up of thematic modules which continually recirculate. The second
movement follows suit, with the new wrinkle that its placid, faintly mournful rst theme returns not only in
its own right but also in a heaven-storming transformation, which recurs as well. The nale embarks on its
own modular cycle, but slowly brings back the music of the primary modules of the two preceding
movements, which it intersperses with its own. The movement—and the symphony—ends when this
process of retrieval is complete.

The resulting sense of time has enormous breadth. The listener is asked for acts of recollection so often that C3.P73
the long span becomes palpable. It expands with no clear borders. As the music proceeds, the force of
recollection makes moving forward in time the same activity as moving backward. The present becomes an
echo chamber of the past. This impression, moreover, arises not only from the moment-to-moment
immersion in the music but also in its overall design. The symphony is a kind of gigantic melodic
palindrome. The rst music to be recollected in the nale (not quoted but recreated; a point on which
Franck insisted) is the melody that begins the second movement. Next comes the music of the module heard
second in the rst movement; last comes the music heard rst. This nal recollection is the most fully
recreated, amounting to an apotheosis of the original melody. For Franck, a church organist all his life, the
elevation of the beginning to form the end is surely suggestive. As the theme ‘slowly piles itself up over [a]
Ninth-Symphony ostinato’, in Donald Tovey’s memorable description (1981: 336), it a ords a further,
symbolic expansion of the timescale for listeners who recall the words of the enthroned gure of Revelations
6:23: ‘It is done. I am Alpha and Omega, the beginning and the end. I will give unto him that is athirst of the
fountain of the water of life freely.’ The music a rms, by enacting, what T. S. Eliot (1962) later wrote of
deep time in the last line of his poem ‘East Coker’: ‘In my end is my beginning’—a palindromic reversal of
the rst line, ‘In my beginning is my end.’

That music of that beginning and end has a famous identity. Its core is a three-note gure paraphrased C3.P74
p. 50 from the nale of Beethoven’s last string quartet, No. 16 in F major Op. 135 (1826). This allusion adds a
novelistic counterpart to the symphony’s epochal time. Beethoven prefaced his nale with an epigraph in
musical type, inscribing the words ‘Muss es sein?’ (Must it be?) under the three-note gure. Franck’s
mournful harmony and gloomy orchestration seem to echo the question in a tragic vein. But whereas
Beethoven (again in the epigraph, anticipating the course of events in the movement proper) answers the
question by inverting the motto over the words ‘Es muss sein!’ (It must be), Franck employs the
palindromic process to transform the question into its own answer. The multiple cross-references of the
symphony thus also occupy the tangled time of the nineteenth-century novel, but, again as in Schubert’s
Symphony No. 9 in Schumann’s reading of it, there are no discernible plots to untangle. The continuous
expansion of the time scale takes the place of the story it would otherwise contain.
The Contracted Scale: Spots of Musical Time C3.S3

The contracted scale echoes the speed of the railway journey, which shrinks distance, induces the uidity of C3.P75
the scene from the carriage window, and requires the very precise measurement of time. (The
standardization of time was a direct result of railway travel (Schivelbusch 2014: 33–44).) The scenic uidity
was especially consequential. At rst widely felt to be disorienting, it eventually shrank the timescale of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
normal perception to an unprecedented degree. Disorientation turned to wonder. As one observer wrote in
1861:

The steam engine […] shifts the point of view every moment; in quick succession it presents the C3.P76
astonished traveler with happy scenes, sad scenes, burlesque interludes, brilliant reworks, all
visions that disappear as soon as they are seen.

(Quoted in Schivelbusch 2014: 61)

After mid-century the felt abruptness with which urban landscapes change, especially in Paris and Vienna, C3.P77
adds an additional in uence. Baudelaire’s poem ‘The Swan’ provides a touchstone: ‘Old Paris is no more (a
city’s form / Changes, alas! more swiftly than a mortal heart).’ Even prior to the 1830s, however, when
European rail transport began to develop in earnest, a tendency toward temporal compression was in
evidence, encouraged by the increasing separation of public and private space (moments snatched for one’s
own use were precious) and the corresponding rise of the lyric in poetry and the art song. The invention of
private time may be one reason why the contracted scale is the special province of the piano—the era’s
primary solo instrument and the crown of domestic space—in contrast to the amplitude demanded for the
public genre of the symphony.

p. 51 The most extreme examples of the contracted scale may be found among some of the earliest: the 24 C3.P78
Preludes for piano Op. 28 that Chopin published in 1839. These Preludes have a claim to being the rst self-
consciously avant-garde pieces in the history of classical music. Chopin composed nothing else like them,
and it remains unclear just what he was trying to do with them; the documentary record is scarce to
nonexistent. Liszt observed that they are not really preludes at all, in the sense that they do not introduce
other pieces. Schumann famously described them as ‘sketches, beginnings of etudes, or, so to speak, ruins,
individual eagle pinions [wing feathers], all disorder and wild confusion’. Those metaphors (there will be
more to say about them) speak to the key recurrent features of the collection (none of which, however, is
consistent; there will be more to say about that, too). Many of the pieces end abruptly or arbitrarily; many
are deliberately ugly by nineteenth-century standards—angular and discordant. Above all, most are very
short. Allowing for the vagaries of performance, roughly ten of them last for less than a minute, and ve of
those for roughly half that. Only one number, the central Prelude, usually known as the ‘Raindrop’, is as
long as a normal short movement or piece, and that one is an outlier, roughly twice as long as the longest of
the other preludes. Whatever Chopin may have been planning, what he did was make a collection that
compresses music to its temporal and expressive limits—limits he would otherwise make a point of
observing.

This pressure extends to the form of the published collection, or, more exactly, to the contradiction between C3.P79
its nominal form and its actual anarchy. On paper, the Preludes constitute a rigorously structured harmonic
cycle that, following the precedent of J. S. Bach, covers all the major and minor keys. Starting with C major
and A minor, the pieces move up the circle of fths pairing each major key with its relative minor. But it is
unlikely that Chopin expected the Preludes to be performed as a complete cycle; he personally never
performed more than three or four as a group. And although the major/minor alternation is obvious in any
full run-through (which has become the concert standard), the harmonic movement is essentially inaudible
except to those with perfect pitch, if they happen to be listening for it. The scheme has little aesthetic
import; it is primarily a means of indexing the pieces.

Various commentators over the years have tried to nd something more in it than that, a mistake I have C3.P80
made myself. But there is nothing to nd, and that is part of the point. The collection is not a cycle, not a
totality, but a miscellany. Schumann’s metaphors are to the point: as ruins, the pieces a ect the ear the way
a picturesque landscape a ects the eye: they charm by intriguing. Most of them (not all, but, again, there is

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
simply no rule that all of the preludes observe) avoid their era’s prevailing aesthetic categories of the
beautiful and the sublime. Instead, as the metaphor of the eagle’s wing feathers suggests, they embody the
residue of high ights that have been missed; they are like the traces of something remarkable that has
passed us by, something we can only surmise. Like individual eagle feathers picked up from the ground, the
preludes have the value found in a souvenir, a talisman, even a fetish. This quality pertains to time as much
as it does to expression: the musical equivalent of a lucky charm is a moment that is almost over as soon as
it begins.

p. 52 In keeping with the spirit of the collection, I will o er only one illustration, in which much more can be C3.P81
found than its duration would suggest. The F minor Prelude is one of the more extreme instances in time
and expression alike. To speak more plainly: Chopin’s Prelude in F minor is a chaotic mess. I say this in
praise of it: the chaos makes sense in its context, which is, precisely, the context of making (or not making)
sense.

The F minor Prelude anatomizes a musical act of violence in a minute or a little less, while also demanding, C3.P82
in the moment before it ends, an act of re ection on what it has just done. The brevity of the piece stems
from something like a desire to silence itself. In the rigidity and single-mindedness of its action it seems to
anticipate the ‘ultimate machine’ devised a little over a century later, in the early 1950s, by Marvin Minsky
at Bell Labs. Once turned on, the machine acts only to turn itself o , a self-cancellation originally
performed by a mechanical hand. But the prelude-machine has some trouble getting itself to turn o by way
of the pianist’s hand—hence it has to take what turns out to be a full minute to terminate itself, a very short
time by normal musical standards, but a strangely long one under the circumstances.

The piece essentially consists of agitated urries of notes followed by increasingly abrupt and forceful C3.P83
attempts to stop them with a single blow. Eight such blows fall and fail in short order. The urries
metamorphose and keep coming; the blows that try to stop them respond in kind. As the piece continues,
the blows break open cracks in the musical surface in the form of notated silences: eighth-note rests
combined with pedal releases after each blow by a heavy two-hand chord. The realization of this damage
requires pianists to observe the pedal indications, which not all will do; failing to observe them loses a good
part of the force of the blows, which is intense in proportion to its futility.

What nally does stop all the commotion is its exile to the bottom of the bass, where it becomes self- C3.P84
consuming in a mass of noise. The noise would have been less drastic acoustically on the pianos of Chopin’s
time than it is on a modern concert grand, but no less drastic expressively. After its racket—a grating trill,
staccato triplets in octaves—what could follow but silence? And silence does. The little eighth-note cracks
abruptly widen and the music splits. For ve quarter-note beats it does not make a sound.

Why so much silence? And what does the player or listener nd there? A void, no doubt: a sense of C3.P85
obliteration or oblivion. Hans von Bülow registered that much when he gave the prelude the ridiculously
literal title of ‘Suicide’, which nowadays still wanders around the internet. But there is much more. Before
we can speculate, however, it is important to recognize that this silence is absolute in one respect and
relative in another. In acoustic reality, the ve silent beats will be lled with background noise, the sound of
people shifting position, coughing, dgeting— even straining to hear. And one way to hear the music at this
point is to hear this silence that rustles with faint sound, eeting though it is. At the same time—literally at
the same time, in what amounts to a kind of two-part counterpoint in hearing—we may hear past the
auditory background and thus hear nothing, in the sense that the prelude as an imaginary rather than an
auditory form blocks out ambient sound. In that dimension the mind’s ear hears the ve beats as silent—
just silent.

p. 53 With what e ect? Does the silence perhaps provide a greater relief from the monomania of the urries than C3.P86
the hammered chords could, a release from the force of compulsion? Perhaps so; and perhaps also a

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
recovery of what passes for normal time, the time that lls the empty beats, itself measured but empty. And
here again there is a test for pianists, many of whom hasten, anxiously, to ll the gap. Perhaps—no, surely
— the silence is a musical articulation of the act of re ection, the internal understanding that follows the
noise of an external event. Perhaps, again, a demonstration of the sheer malleability of the time sense,
articulated in the di erence between the striking visual emptiness in the score, composed of the white space
around the half-note rests in sharp contrast to the calligraphic elaborateness of the preceding notation, and
the actual duration, in seconds, of the silence.

That duration is ambiguously both long and short. The silence lasts just one beat longer than the cadential C3.P87
half-note chords that follow lamely; this is a false ending. The cadence sounds like a lie the music is telling
itself, putting a good face on its failure to do anything more than stop short. But perhaps above all the
duration of the silence is the time of the same uncanniness that occurred to Arthur C. Clarke (1959: 159)
when he observed the ultimate machine built by Claude Shannon after Minsky’s model:

Nothing could be simpler. It is merely a small wooden casket, the size and shape of a cigar box, C3.P88
with a single switch on one face.

When you throw the switch, there is an angry, purposeful buzzing. The lid slowly rises, and from C3.P89
beneath it emerges a hand. The hand reaches down, turns the switch o and retreats into the box.
With the nality of a closing co n, the lid snaps shut, the buzzing ceases and peace reigns once
more.

The psychological e ect, if you do not know what to expect, is devastating. There is something C3.P90
unspeakably sinister about a machine that does nothing—absolutely nothing —except switch
itself o .

As a description of the F-minor Prelude, that is not bad at all. C3.P91

The Living End C3.S4

Clarke’s use of the phrase ‘wooden casket’ overtly links the ‘unspeakably sinister’ to mortality, as if the C3.P92
machine had been programmed to deny the value of life. (Von Bülow’s ‘Suicide’ applies to the device better
than it does to the music.) The problem is not death itself but death as a short circuit. Some further insight
into this feeling, and into Chopin’s prelude, can be drawn from an unfashionable source. In Beyond the
Pleasure Principle (1961), Freud notoriously proposed the existence of a death drive. The concept almost
immediately met with dismissal, even contempt, yet it has had a remarkably long, well, life. One of Freud’s
key questions was why, if the death drive exists, the organism should bother to live at all? His answer,
p. 54 famously, is that life is a detour on the way to death; the life drives are recruited to delay death so that
each organism can die in its own way (p. 47). Life, in e ect, assumes its value as a way of giving death
meaning.

Precisely that meaning is what the ultimate machine refuses. And something like that meaning is what C3.P93
Chopin’s prelude fails—fails deliberately—to achieve. The prelude adopts the shortest possible timespan
between switching on and switching o . Because the music is expressive rather than mechanical, it must do
a little more than ‘absolutely nothing’ before it cancels itself. It ventures the detour to demonstrate that no
detour is possible. Instead of mechanism, it intimates nihilism. And then it shuts itself up and turns itself
o .

Chopin approaches the borders of nihilism in other very short pieces too, most notably in the nale of his C3.P94
Piano Sonata No. 2 in B♭ minor Op. 35 (1839), whose swift, faint octaves do nothing, absolutely nothing, to

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
counteract the force of the famous funeral march heard earlier. But in the present context the most
signi cant thing about the prelude, and others in the collection, is not its particular expressive force, but
that fact that a minute or less has become a viable medium for an expressive totality.

Note
1. Translations from Schumann, and from Nietzsche and Baudelaire, are mine. C3.N1
References C3.S5

Baudelaire, C. (1861 [2004]). Les Fleurs du mal. https://fleursdumal.org/. C3.P95


Google Scholar Google Preview WorldCat COPAC

Clarke, A. C. (1959). Voice across the sea. Harper & Row. C3.P96
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
Derrida, J. (1980). The law of genre (trans. A. Ronell). Critical Inquiry 7: 55–81. C3.P97
Google Scholar WorldCat

Eliot, T. S. (1962). Collected poems, 1909–1962. Harcourt, Brace & World. C3.P98
Google Scholar Google Preview WorldCat COPAC

Freud, S. (1961). Beyond the pleasure principle (trans. J. Strachey). Norton. C3.P99
Google Scholar Google Preview WorldCat COPAC

Lyell, C. (1830–33). Principles of geology: Being an attempt to explain the former changes in the earthʼs surface. C3.P100
https://books.google.com/books?
id=JoW8AAAAIAAJ&dq=intitle%3APrinciples%20intitle%3Aof%20intitle%3AGeology%20inauthor%3ALyell&pg=PR4#v=onepage
&q&f=false
Google Scholar Google Preview WorldCat COPAC

McPhee, J. (1982). Basin and range: Annals of the former world. Farrar, Straus and Giroux. C3.P101
Google Scholar Google Preview WorldCat COPAC

Nietzsche, F. (1895). Also sprach Zarathustra. C. G. Naumann. https://books.google.com/books? C3.P102


id=DDMPAAAAYAAJ&dq=intitle%3Aalso%20intitle%3Asprach%20intitle%3Azarathustra&pg=PP9#v=onepage&q&f=false
Google Scholar Google Preview WorldCat COPAC

Playfair, J. (1822). Biographical account of the late James Hutton, M.D. In J. G. Playfair (ed.), The works of John Playfair, vol. 4. C3.P103
Constable. https://books.google.com/books?
id=j9MyAQAAMAAJ&dq=intitle%3ABiographical%20intitle%3AAccount%20inauthor%3APlayfair&pg=PP13#v=onepage&q&f=fals
e
Google Scholar Google Preview WorldCat COPAC

Rancière, J. (2006). The politics of aesthetics (trans. G. Rockhill). Continuum. C3.P104


Google Scholar Google Preview WorldCat COPAC

Schivelbusch, W. (2014). The railway journey: The industrialization of time and space in the nineteenth century. University of C3.P105
California Press.
Google Scholar Google Preview WorldCat COPAC

Schumann, R. (1854). Gesammelte schri en über musik und musiker. Dritter band. G. Wigand. https://books.google.com/books? C3.P106
id=jfxaAAAAQAAJ&dq=editions%3Amr-BYUVD0yEC&pg=PP5#v=onepage&q&f=false

p. 55 Stendhal (1989). The red and the black (trans. S. Haig). Cambridge University Press. C3.P107
Google Scholar Google Preview WorldCat COPAC

Tennyson, Alfred Lord (1850). In memoriam A.H.H. http://www.online-literature.com/tennyson/718/. C3.P108


Google Scholar Google Preview WorldCat COPAC

Tovey, D. (1981). Essays in musical analysis: Symphonies and other orchestral works. Oxford University Press. C3.P109
Google Scholar Google Preview WorldCat COPAC
Whitman, W. (2001). Leaves of grass and other writings (ed. M. Moon). Norton. C3.P110
Google Scholar Google Preview WorldCat COPAC

p. 56 Wordsworth, W. (1979). The Prelude: 1799, 1805, 1850 (ed. J. Wordsworth). Norton. C3.P111
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353468984 by National Science & Technology Library user on 26 May 2023
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
CHAPTER

4 Music as Time, Music as Timeless  C4

Kristina Knowles

https://doi.org/10.1093/oxfordhb/9780190947279.013.3 Pages 57–76


Published: 08 December 2021

Abstract
This chapter presents a framework for parsing di ering conceptual and analytical positions on time in
music, focusing speci cally on two contrasting ideologies. The rst perspective views music as an art
form that exists only in and through the unfolding of time; the second views music as capable of
evoking a static temporality, referred to by many scholars as a sense of stasis or timelessness.
Discussions on the relationship between time and music typically engage with a subset of overlapping
and interacting positions on time. Time is sometimes analysed as external and objective, but can also
be construed as internal and subjective. Finally, time is also understood to be created or represented by
music, an idea encapsulated by the term ‘musical time’. References to timelessness in music engage
with these latter two views on time, speci cally music’s ability to represent temporal concepts
associated with speci c structures (musical time) and perceptual mechanisms related to certain
musical features that result in a subjective experience interpreted as timelessness. Using the dual
lenses of psychology and philosophy, I argue that timelessness is an inherent part of the multiple
systems of temporal production and perception that underlie the way we experience and discuss time
in music.

Keywords: music, time, timelessness, stasis, perception


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

FOR many scholars, the idea that music not only occupies but shapes time is a fairly easy one to accept. In C4.P1
music—or, perhaps, because of music—time is understood to do all sorts of things: to accelerate (Epstein
1981), to slow down (Clifton 1983), to reverse (Gri ths 1985), to split into multiple concurrent timelines
(Kramer 1988; Hyland 2016), to become cyclic (Berger 2007; Clayton 2000), to become linear (McClary
2012), to become vertical (Kramer 1988), to suspend time (Klein 2004; Margulis 2014; Taylor 2016), to stop
time (Frith 1996; Thurmaier 2006). These last two touch on a less common, more contentious, and yet still
existent idea: that music can be timeless. Such a notion can be found peeking through the work of various
scholars, from Basil de Salincourt’s assertion that music ‘suspends ordinary time’ (Langer 1953: 110),
Raymond Monelle’s suggestion that ‘music invites us to step out of time into its own timeless state’ (2000:
86), or Diana Luchese’s psychologically grounded claims that a perception of changelessness in Messiaen’s
music leads to an experience of ‘timelessness’ (2010).

What is meant by ‘temporal suspension’ or ‘timelessness’ here is not always the same. For some, the notion C4.P2
of timelessness in music is purely conceptual (Hasty 1997), while others would argue that it is a perception
or experience a orded by speci c characteristics associated with a subset of compositions (Frith 1996;

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
Kramer 1988; Luchese 2010). Other scholars oppose the concept of timelessness in music outright, citing the
1
oft-held belief that music exists because time passes, that music is inherently temporal (Clifton 1983;
Epstein 1981), and therefore timeful, not timeless. Yet underlying all these various claims is a shared belief
that music interacts with and (for some) plays with the medium through which it sounds.

The rather reductive dichotomy implied by the title of this chapter is intended to capture these opposing C4.P3
2
views on music’s ability to evoke (conceptually or perceptually) notions of timelessness in music. As with
most dichotomies, much is lost in categorizing arguments and positions on music’s temporality into one
p. 58 camp over another, and various nuanced and insightful perspectives may be left out in the cold. Instead
of arguing for one position or the other, I will explore timelessness as an inherent part of the multiple
systems of temporal production and perception that underlie the way we experience and discuss time in
music. In doing so, I will argue that the tension between views of music as inherently expressing the ow of
time and the belief that music is capable of evoking ‘static temporality’ arise out of di ering but
intersecting notions of where and how time is constituted in music. Understanding the ways in which music
scholars conceive, construct, and engage with the relationship between music and time may shed light on
the origins of these di ering ontological claims, clarify the di erent ways in which music and time interact
and are co-created, and provide a framework for future research.

A Multitude of Time(s) C4.S1

Investigations of the interaction between time and music are commonly plagued by the metaphysical C4.P4
complexity of time itself, often leading to convoluted questions concerning the nature of time. Rather than
attempt a Sisyphean feat, I shall sidestep the question of time’s metaphysical nature and instead suggest
that we as human beings experience a number of di erent ‘times’. These can be organized in a variety of
ways, such as types of time that are environmentally based (solar time, lunar time, geological time,
changing of the seasons), biologically based (circadian rhythms, menstrual cycle, timing of physical
3
gestures), or culturally based (religious time, social time). Notions of time in discussions of music often
derive from an assumed binary between two types of time: the rst being measurable, discrete, and
communal; the second subjective, uid, and individual (see Table 4.1). A similar distinction can be seen in
the eld of psychology, which distinguishes between objective clock time and subjective psychological time.
To a limited extent, these two types can be mapped onto di erences between Newtonian concepts of time as
equably owing versus Einsteinian relativist notions of time as individually constituted.

Despite the similarities between a view of time as measurable and a Newtonian concept of time, I do not C4.P5
4
wish to invoke the concept of an absolute time equably owing here. Rather, the use of the term
‘measurable’ is meant to highlight a particular way of approaching or understanding time, and the unit by
which one measures time in and of itself is less relevant than the act of treating time as measurable. The
value of approaching time in this manner lies in its ability to enable collaboration between individuals—it
5
serves a social and thus communal purpose. While this can be seen in the everyday use of calendars and
clocks to make and keep appointments, or more communal and sonic markers of time such as church bells
and clocktowers, it is also present in the ‘5–6–7–8’ count-o used by dancers to coordinate movements
(Short 2019) and the gestures of a conductor as she marks tempo and metric beats for an ensemble. In all of
p. 59 these instances time is demarcated into meaningful units, its measurement supporting and enabling
interpersonal coordination. In contrast, the second type of time—let us call it ‘internal time’—is not the
time of the group but of the individual; it is time as experienced, time as lived, time as shaped by perception
and event. It is, arguably, the time with which most musicologists and psychologists concern themselves,
the former for the possibilities this perspective a ords when considering music as constructing its own
temporal experience, and the latter for the possibility of arriving at a better understanding of states of
consciousness. Most notably in music scholarship, the event itself can have its own time, a notion

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
encapsulated in the frequently used and rarely de ned concept of ‘musical time’. This third, event-based,
music-created time is subsumed within the second type of time in Table 4.1. However, its use and treatment
within the literature suggests that it may be worthwhile to consider it a third type, one that interacts and
intersects with the other two.

Table 4.1 Terms used by scholars for Type 1 and Type 2 Time C4.T1

Type 1 Type 2
measurable, discrete, communal subjective, fluid, individual

ʻordinary timeʼ ʻmusical timeʼ

(Kramer 1988) (Kramer 1988)

ʻobjective timeʼ ʻsubjective timeʼ

(Epstein 1981; Delaere 2009) (Epstein 1981; Leong 2001)

ʻcommon-sense timeʼ ʻlivedʼ or ʻexperiencedʼ time

(Langer 1953) (Langer 1953; Monelle 2000)

ʻontological timeʼ ʻpsychological timeʼ

(Reiner 2000) (Reiner 2000)

The notion that music may intersect with multiple di erent ways of viewing and understanding time is not, C4.P6
in and of itself, new. The three categories proposed above can be found within Thomas Clifton’s observation
that ‘there is a distinction between the time which a piece takes and the time which a piece represents or
evokes’ (1983: 81); as demonstrated in Figure 4.1. Clifton’s consideration of the time music takes falls into
the category of external time and belongs most properly to the subtype of clock time as it entails the
duration of the musical work. Conversely, the time which a piece represents is associated with the concept
of musical time, itself involving compositional constructs of time as well as considerations of the temporal
signi cance assigned to certain musical gestures in di erent time periods and across di erent cultures.
Finally, the time which a piece evokes suggests the experience of the individual and the uid, subjective
temporality encapsulated by the notion of internal time discussed above.

Let us return, then, to this concept of musical time and interrogate what it means for music to represent C4.P7
time. From the start, the idea of music representing or communicating a temporal concept nds resonance
p. 60 with semiotics and brings to mind Raymond Monelle’s assertion that ‘music can also signify time. There
is a temporality of the signi ed as well as a temporality of the signi er […] Music, then, can subsist in time
without taking time’ (2000: 88). Arguably the process of semiosis is inherently involved in the evocation of
a musical time, as the two terms that make up this concept are themselves involved in semiological
processes: time because of the inherent di culties surrounding its nature, which is best understood
through various things that act as signs of time (i.e. clocks, duration, rhythm, continuity, etc.) and music
which, in terms both of an individual work and of the art form as a whole, functions as a system of signs
(Reiner 2000). As a result, musical time participates in a semiotic network with the individual concepts of
both music and time, and by extension with the network of semiological meaning for which each of these
concepts is at the centre. To put it another way, concepts pertaining to musical time can encompass both
discussions of temporal qualities that are attributed to music, and music’s potential to represent concepts of
musical time, of which the listener may then become aware. Understanding musical time in this way allows
for both the acknowledgment and nuancing of di erent perspectives within scholarship on time and music

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
that can be represented by Figure 4.1.

Figure 4.1

A Venn diagram depicting the di erent interacting concepts of time that can be applied to discussions of music. C4.F1

Out of this framework comes the assertion that musical time can be associated with (1) the duration of a C4.P8
musical work, (2) compositional structures, processes, procedures, and concepts which pertain to the
temporal structure of music, and (3) the subjective experience of music and temporality on the part of the
listener. While there is scholarship premised on this rst notion of musical time, such as Kramer’s
emphasis on proportional listening in the music of Stravinsky or the numerous articles on the Golden Ratio
6
in the works of various (predominantly twentieth-century) composers, most scholarship falls into these
latter two camps and thus they will be the focus of the remainder of this chapter.

p. 61
Musical Time: Lyric Time C4.S2

An understanding of musical time which encompasses music’s capacity to represent temporal concepts that C4.P9
may also be perceptible to a listener can be found in Raymond Monelle’s distinction between lyric time and
progressive time (2000). A number of scholars have described the temporal qualities of the former as stasis
or timelessness (Klein 2004; Monelle 2000), while the latter is understood to involve a dynamic and
processive type of time. A similar distinction is made by Robert Hatten between Satz and Gang, building on
work by A. B. Marx and suggesting that the former is present-oriented while the latter is future-oriented
7
(1997). In both cases a contrast is made between a lyrical theme with clear melodic emphasis, harmonic
stability, and phrase repetition implying a temporality of stasis or—perhaps more accurately, an expanded
present—and a transition or development functioning passage which contains more rapid guration,
harmonic motion, sequences, and modulation. These elements combine to create a sense of temporal
progression and dynamic motion.
In many ways, lyric and progressive time (Satz and Gang for Hatten) can be mapped onto the categories of C4.P10
Being and Becoming respectively, as these oft-cited temporal concepts represent the type of time these
di erent musical passages aim to convey. From here, a loose connection can be drawn with Jonathan
Kramer’s two primary temporalities which he calls ‘linearity’ and ‘nonlinearity’—categories which he
equates with becoming and being respectively (1988: 16). Though both sets of scholars utilize similar
underlying temporal concepts which are connected to speci c types of compositional choices, there are key
di erences. In Kramer’s work, the type of musical time expressive of becoming is predicated on notions of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
causality; initial events imply later ones and later events can be understood as consequences of earlier ones.
In the work of Hatten, Klein, and Monelle, the notion of progressive temporality is compositionally located
within speci c sections of pieces that carry set formal functions (namely transitional and developmental
sections) and, for Monelle in particular, is part of a larger cultural dichotomy of temporal concepts relevant
to the Romantic era.

Nonlinearity, the musical time of being, applies to pieces that are governed by a principle or set of C4.P11
tendencies. According to Kramer, ‘nonlinearity is a concept, a compositional attitude, and a listening
strategy that concerns itself with the permanence of music: with aspects of a piece that do not change, and,
in extreme cases with compositions that do not change’ (1988: 19). In contrast, lyric time carries very
speci c connotations in the Romantic era, not just an extended present or stasis but a ‘lostness’, a kind of
‘space where the remembered and imagined past is re ected’. Such a space is critically contrasted with the
mobility a orded by progressive time, which Monelle suggests ‘is a forum for individual choice and action
that is ultimately doomed’ (2000: 115). This description detailing the function of lyric time within the
Romantic period resonates with the temporality of the pastoral, which also carries connotations of
p. 62 timelessness and a sense of nostalgia (Leydon 2003; Marinelli 1971; Peattie 2002). It is no coincidence,
then, that many expressions of lyric time within the nineteenth century are also instantiations of the
pastoral topic, with its emphasis on tonic and subdominant harmonies, simple melodies, lilting rhythms,
and repetition. Such musical elements easily map onto Klein’s description of lyric time as ‘those
presentational sections in which melody comes to the fore, and in which harmonic and phrase structures
are relatively stable’ (2004: 37).

The opening of Chopin’s Ballade No. 2 in F major Op. 38 (1839) portrays both lyric time and the pastoral C4.P12
topic (Figure 4.2). These two types of musical time—one direct, the other a musical topic with strong
temporal connotations—combine to provide a nuanced temporal expression whose meaning gains depth
and dimension as the piece unfolds. The lilting rhythm of the siciliano begins with a drone and then widens
out to an F-major melody, its opening 10 bars featuring a gentle tune that alternates with what Jonathan
Bellman terms a ‘bagpipe-like choral response’ (2010: 16). The emphasis on subdominant harmonies,
relatively static harmonic and phrase structures, and repetition of the opening material signi es both the
nostalgic dream of the pastoral and the endless present of lyric time, the two creating a kind of timeless
bucolic ideal. The subtle shading of loss that always hazes the edges of the pastoral peeks through the
contrasting phrase of this opening salon piece, it’s tonicization of A minor foreshadowing the coming
storm. As implied in the Monelle quotations above, passages expressing lyric and progressive time are often
juxtaposed, the latter used to provide the harmonic and structural movement that the former typically
avoids. In this instance, the opening temporality is e ectively and violently shattered by a stormy A-minor
brilliant topic reminiscent of Chopin’s Étude Op. 25 No. 11 in A minor, ‘Winter’ (1836). The interruption not
only serves to provide motion, contrast, and harmonic movement but provides a narrative force as well.

Subsequent returns of the pastoral topic and its invocation of lyric time are progressively coloured by the C4.P13
instability of the A-minor theme and its key, not only amplifying the sense of loss and nostalgia that lurked
in the opening bars but creating an ever-increasing sense of temporal distance. The length, repetition, and
harmonic stability of the opening passage may be su cient for some listeners not only to grasp the musical
time portrayed but also to experience an expanded moment, a sense of suspended present. While
subsequent iterations of the pastoral theme may recall this sensation, they are neither long enough, stable
enough, nor repetitive enough to recapture that temporal experience, thus adding to the layers of
signi cation and temporal meaning the theme has accrued by its nal appearance, rmly in the key of A
minor. Such a reading is not only strengthened by the stark contrast in key, rhythm, guration, harmony,
melodic emphasis, and phrasing between the F-major pastoral and the A minor storm, but also the absence
of any attempt to transition between the temporality of the two.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
In other, less common cases, composers have written entire pieces expressing lyric time. Schumann’s C4.P14
‘Träumerei’ from Kinderszenen Op. 15 (1838) is one such example, its title itself implying the kind of dreamy,
timeless quality it seeks to convey. While this piece contains succession, rhythm, metre, repetition,
p. 63 implication, and resolution, every cadence—rather than leading to the completion of an idea or building
upon one another to contribute to the articulation of a larger and more formally signi cant moment of
arrival—simply melts into another statement of the melody. The music, as a whole, fails to go anywhere,
instead representing or, for the attuned listener, evoking an experience of an endless present.

Figure 4.2

Excerpts from Chopinʼs Ballade No. 2 in F major Op. 38 (1839). Excerpts (a) and (b) show passages from the opening siciliano C4.F2
theme and the foreshadowing of A minor. Excerpt (c) is the opening of the A-minor theme, and Excerpt (d) shows the return of
the siciliano theme at the end of the piece, now transformed to the key of A minor.

p. 64 In the cases of lyric time discussed here, the listener themselves may or may not have a change in their C4.P15
subjective experience of time, what Jason Noble (2017) calls an ‘induction of timelessness’, though the
possibility is certainly there. And, it is worth noting, if the listener is unfamiliar with the musical style of
this period, they may also be unaware of the type of musical time being communicated. As Barbara Barry
notes, ‘the experience of musical time is very personal and subjective’ (1990: 228). Within the semiological
tripartite framework of poiesis, trace, and esthesis presented by Thomas Reiner (adopted from Jean Jacques
8
Nattiez), it is ultimately within the process of esthesis that musical time resides. Though the composer may
intend to represent a particular concept of time within the musical work (in this case, lyric time), the
listener may not perceive within the sonic trace of the piece the intended temporal concept. ‘If a listener is

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
not aware of musical time, there is no reason to assume that this listener experiences musical time’ (Reiner
2000: 66). Alternatively, the listener may be aware of time without associating this awareness with the
music to which they are listening, and thus there is no musical time. This is not to say that types of musical
time which are not widely perceptible are not worthy of discussion, but rather that there can be a disconnect
between notions of musical time that are meant to be heard and experienced and those that are intended to
describe particular facets of musical structure and organization. While this latter group does impact what is
heard, they may not activate notions of time or temporality in the mind of the listener.

To return to the question of lyric time, the perception and understanding of this type of musical time C4.P16
requires a double signi cation, wherein the listener (or the analyst!) moves from a perception and
recognition of certain musical features to an interpretation of musical time (lyric time), and from the
interpretation of musical time to the concept of time itself in all of its varying forms, including concepts
such as the endless present, timelessness, and stasis. The speci c qualities evoked by lyric time within the
nineteenth century are located in this second process of signi cation, deepening and nuancing the broader
temporal concepts of an extended present, tinging it with the sense of loss and longing that in ects many
Romantic works.

As the proceeding examples demonstrate, concepts of musical time can revolve around compositional C4.P17
processes and formal structures and may involve cultural concepts which shade the temporal meaning of
the passage in question. Within the chain of signi cation described above, musical time acts as an
intermediary node between a listener’s perception of music and their awareness or conceptualization of
time in response to the music with which they engage. In the words of Jonathan Kramer, though the time
music creates (‘musical time’) and the time the individual experiences (‘ordinary time’) may ‘lead parallel
lives’, it is possible for one to listen so deeply as to lose the distinction between the two (1988: 7). A similar
sentiment is echoed by Reiner, who suggests that a listener may either be aware of their individual sense of
time (‘psychological time’) and establish a conceptual link between it and the music or may think that the
music is determining their sense of time and thus lose the distinction between the two (2000: 136). Such an
idea moves beyond the middle stage of the semiological process outlined above in which speci c concepts or
p. 65 types of musical time mediate between the musical features and the listener’s experience of time, and
suggests a direct connection between the perception of music and an individual’s perception or conception
of time.
Experiencing Timelessness C4.S3

The concept of internal time privileges both the individual’s perception of time and their interpretation of C4.P18
that perception. When intersecting with musical time, the focus shifts to address how the structure and
pacing of the music in uences one’s subjective experience of time and leaves open the possibility that such
an experience may be understood, conceived, and thus in a way perceived as timelessness, stasis, or an

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
endless present. In such instances, musical time and the subjective experience are not grounded in the
external, the measurable, or the discrete but rather within Kant’s notion that time is individually
9
constituted. Psychologists have demonstrated that experienced events can manipulate the rate at which
individuals perceive time to pass, in a sense creating an experience of temporal acceleration or deceleration
unique to each particular individual. If we accept that such concepts are not just plausible but widely
accessible to the general population (as implied in popular colloquial sayings such as ‘Time ies when
you’re having fun’ or ‘That class dragged on for forever’), then it is only a small step to suggest that an
individual’s perception of an event may evoke concepts of timelessness should the temporal components of
the event itself exceed some norm or signify the related temporal concept.

In general, most scholars are comfortable with the idea that music can be fast or slow, can accelerate or C4.P19
decelerate, and that there can be di erences between a perceived tempo and a notated tempo such that we
hear the pacing of the music to be slower than is indicated in the score based on the metrical level to which
10
we entrain. These are all readily acceptable temporal constructs in part because they map nicely onto
notions of motion and movement, resurrecting Aristotle’s belief that time is motion (and, it is worth noting,
subject to the same aws). However, other temporal concepts such as time stops, time suspended, endless
present, or timelessness seem more problematic, less grounded in a world of physics and motion. These
temporal terms, which we might broadly lump under the category of ‘out of time’ (though it is worth noting
that in some of these, time is still present and moving although its motion has been demoted in importance
or awareness), do not imply that from an external perspective time has stopped. Rather, they are attempts
to capture an experience, a feeling, the colour and quality of a particular moment. They are, in other words,
more cultural and conceptual constructs than physical realities, but that no more lessens their relevance or
11
impact than does any other culturally grounded idea.

As suggested above, one way of evoking a perception of timelessness is for temporal components to exceed C4.P20
some kind of norm. In the case of music, Jason Noble has suggested that the norm exceeded is ‘human
time’, which he describes as temporal scales and structures which align with ‘optimal human auditory
p. 66 information processing and embodiment’ (2017: 5). His focus on aspects of rhythm and metre such as
tempo, beat rate, and metric complexity in discussions of timelessness in music nds resonance with
scholarship on Messiaen which aims to highlight how speci c musical features may evoke a perception or
experience of timelessness on the part of the listener. Note that a distinction is being drawn between the
perception of timelessness as a concept being portrayed or represented by the music and the subsequent
induction of timelessness for the listener. Arguably the former is a more common occurrence than the
latter, and falls more into categories of musical time than internal time. However, the latter represents a
valid, if less frequent, range of experiences a orded by music. In suggesting that music may induce an
experience of timelessness for the listener, it is worth reiterating that this is a perspective grounded in the
ontological source of time as internal—that is, as the time that exists for the individual and not for the
12
collective. A group of people listening to the same piece may have very di erent perceptions and
experiences—while one may perceive the music communicating concepts of timelessness, another may
have an experience that they describe as timeless and a third may neither perceive nor experience
13
timelessness while listening.

Returning to work on Messiaen, a number of scholars have explored how the composer evokes timelessness C4.P21
or eternity as part of the spiritual message of his music, focusing on both symbolic musical elements that
Messiaen associates with eternity as well as sonic characteristics of the music that create the perceptual
illusion of timelessness or stasis. Features consistently evoked in this body of work in association with
evocations of timelessness in the music include the lack of metre, extremely slow tempo, deceleration,
silence, and static textures or harmonies. In particular, Paul Gri ths describes the lack of metre in
Messiaen’s music as a means by which the past, present, and future are blended together. ‘Instead of metre,
which gives each moment in a bar a di erent signi cance and hence fosters a sense of orderly progression,
Messiaen’s music is most frequently tied to a pulse, which insists that all moments are the same’ (1985: 15).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
For Diane Luchese, the absence of a metre contributes to a reduction in the level of dissonance in Messiaen’s
music, as potential con icts between metre and rhythm (metric shifts, displacement, syncopation) are
eliminated. This minimization of contextual dissonance is one of the features that she argues contributes to
a sense of timelessness.

It is important to note that Luchese grounds her arguments about eternity in Messiaen in the notion that an C4.P22
impression of changelessness in music can then suggest to the listener the notion of timelessness. The
majority of her arguments are grounded in perception and are informed, to varying extents, by psychology.
Through the use of low rhythmic density, long durations between pulses, slow rates of harmonic change,
and silence, Luchese argues that Messiaen ‘stretches out information over time, to manipulate our
perception of motion and duration, and thereby create an impression of changelessness’ (2010: 182;
emphasis original). The movement entitled ‘Louange à l’éternité de Jésus’ from Messaien’s Quatuor pour la
n du temps (1941) o ers an example of many of these characteristics. The direction to perform the piece
In niment lent, extatique (in nitely slow, ecstatic) combined with the indicated tempo of sixteenth note =
p. 67 c.44 bpm stretches the perceptual ability of the listener to perceive a coherent melodic line, a gestalten,
rather than a series of consecutive notes. As seen in Figure 4.3, the rst note of the cello solo lasts for c.2.8
seconds. While the following sixteenth notes provide a brief sense of motion that may help the listener hear
this opening gesture as a uni ed whole (the entire three-note motive taking c.5.6 seconds to perform), the
following series of eighth notes, each lasting approximately 2.8 seconds, pushes against perceptual
boundaries. Within sparse musical environments made up of singular sounding events, durations that
exceed 2 seconds quickly become mensurally indeterminate (Hasty 1997). In other words, we lose the ability
to reproduce, and thus understand and experience, the rhythmic relationship between singular events that
last longer than 2 seconds.

Figure 4.3

A duration analysis of the opening for the 5th movement from Messiaenʼs Quatuor pour la fin du temps (1941). Durations are C4.F3
based on the indicated tempo in the score.

Furthermore, the limitations of the perceptual present, understood to encompass a duration of C4.P23
14
approximately 5–6 seconds makes it harder for a listener to grasp the melodic line as it unfolds. The nal
quarter note of the melody sustains the pitch E5 for c.5.6 seconds, creating the potential for this one
sustained tone to completely ll our present awareness, generating the perception of a frozen moment,
15
suspended in an endless wait for change. The steady pulse initiated by the piano midway through this note
provides that change, though the duration of each chord is still quite slow (each sixteenth note lasting
approximately 1.4 seconds). The impression of timelessness evoked by the opening of ‘Louange à l’éternité
de Jésus’ colours the music that follows, having established the primary temporal quality for the
16
movement.

Though the entrance of the pulsed chords may break the spell—so to speak—for some listeners, the C4.P24
impression of timelessness which initiates this movement inevitably colours the passages that follow. It is

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
worth noting that not every passage within a piece needs to be perceived or conceived of as ‘timeless’ for the
concept to be relevant. A distinction can be drawn between the macrocosmic temporal meaning of a work
and the microcosmic uctuation of temporal experience, and notions of timelessness can be present at
either level.

Many of the musical features or characteristics listed here as potentially evoking an experience of C4.P25
p. 68 timelessness for the listening subject can be related to broader psychological mechanisms concerning
in uences on our subjective experience of time. While the literature on time perception in psychology is
complex and multifaceted, a few key ideas are helpful in considering musical contexts that may evoke a
perception of timelessness. Speci cally, complexity of the content and the nature and degree of one’s
attention towards aspects of the duration or event have been shown to in uence the encoding and
judgement of time. In what follows I will brie y describe each of these components as they are discussed in
the eld of psychology, and explicate their possible application to music.

The idea that the number of events occurring within a set duration can a ect one’s judgement of that C4.P26
duration has a fairly long history in psychology. Early instantiations of this concept can be found in the
works of Paul Fraisse (1963) and Robert Ornstein (1969), and are based around information retrieval
theories that draw parallels between computer processing and human processing. In short, the more
information within a duration, the more content to be encoded into memory, and thus the longer it will take
to recall this information, leading to a retrospective association of temporal expansion with lled durations
in comparison to empty durations. While these earlier models are based on static structures predicated
around how much information one can store in memory, later theories introduced more dynamic aspects of
content, including the number of perceived changes within an interval of time (Block and Reed 1978) as well
as di erences between experiencing a duration in the moment versus a retrospective recollection of a
duration (Block and Zakay 1997; Block et al. 2010). This former point suggests that when considering the
temporal e ects of a musical passage, shifts in texture, register, mode, dynamics, metre, and other
elements can be relevant factors in the construction of a temporal experience. As Luchese notes, a
perception of changelessness may prompt a perception of timelessness. In regard to this latter point, when
a duration of time is experienced in the present, a lled duration is perceived as shorter than an empty
duration of the same absolute length—in other words, time moves faster with a lled duration and slower
with an empty duration (Block 1985; Ihle and Wilsoncroft 1983). Such ndings provide empirical support for
the commonly experienced di erence between 15 minutes spent sitting in a waiting room and 15 minutes
17
spent engaged in activity.

In the case of music, complexity is often a better parameter than density, as musical passages can contain a C4.P27
high number of sounding events within a thick texture but not be perceived as complex when one considers
the harmonic content or durational rhythms of the passage. Rather than considering the pure number of
events within the music (an issue of density), it is better to consider the degree to which events are
comprehensible (an issue of complexity). Given the ndings from psychology discussed above, during the
process of listening to a musical passage, higher complexity can result in an experience of temporal
contraction while lower complexity can result in an experience of temporal dilation. However, the
relationship between complexity and temporal perception is not collinear. Highly complex musical passages
can result in overstimulation, overwhelming listeners with more layers, textures, and events than they can
p. 69 meaningfully parse. In cases where musical complexity exceeds the cognitive capacities of the listener,
the temporal experience evoked is akin to that of a sparse musical passage—timelessness or stasis. In the
words of psychologist Wayne Hogan, ‘time that is too lled may a ect the individual’s perception of
duration in much the same way as it is a ected by empty time […] sensory deprivation (underestimation)
and sensory overload (overstimulation) may be viewed, then, not as polar opposites, but rather as two
expressions of the same phenomenon’ (1975: 33). Musical characteristics such as low rhythmic density,
slow tempo, a sparse texture, and infrequent or slow harmonic changes can prompt experiences of
timelessness during listening, as can extremely fast tempos and pulses, erratic rhythms, rapid harmonic

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
changes, and aleatoric textures.

The second psychological mechanism discussed here, the question of attention, involves both what the C4.P28
individual is attending to—temporal information or non-temporal information—and the temporal
structure of the stimuli in question (Jones and Boltz 1989; Zakay 1989). Theories of attention in time
perception function on the understanding that humans have limited attentional resources that are ‘divided
amongst all the tasks that the organism must perform’ (Zakay 1989: 371). When attentional resources are
devoted more towards the processing of non-temporal information, loosely de ned as a stimulus or activity
that distracts from an awareness of time, our perception of time speeds up and we ‘lose track’ of time. In
contrast, when attentional resources are directed towards the processing of temporal information in the
form of awareness of duration or time, our perception of time slows down; time ‘drags’. Applications of this
theory are complicated, as music is broadly considered to be an inherently temporal art form, and thus all
components of a musical stimulus could be construed as temporal in some sense. There are, however, a few
clear instances where this theory might apply. When in the context of silence, we often nd ourselves in the
condition of waiting for something, either for the end of the piece or the next note. This ‘waiting’ prompts
18
an attention towards the duration of the silence and thus an expansion in our subjective sense of time.
Arguably, a sustained tone could evoke a similar perception. Beyond this, moments where music is not
signifying temporality for the listener may be construed as instances of non-temporal information in
relation to music, and some musical elements may prompt an awareness of, or focus on, time more than
others.

The elements of rhythm and metre in particular may be relevant here, as they impact the temporal structure C4.P29
of the event itself, another factor in attentional models. Speci cally, highly coherent temporally structured
events that are hierarchically organized—in other words, metrical passages—allow for future-oriented
attending. In contrast, less coherent events encourage analytic attention, requiring people to attend to
relationships among adjacent items within a duration (Boltz 1989; Jones and Boltz 1989). Applying these
concepts to music, passages where the metre is clearly articulated across multiple hierarchical levels require
less attention than passages that are either metrically ambiguous or non-metred, as they a ord a more
future-oriented mode of attending that allows for the generation of expectations. Unmetred or metrically
ambiguous passages of music, which are unpredictable, are more likely to prompt a focus on the unfolding
‘now’, and this increased attention to a narrow duration can result in a slower subjective sense of time
p. 70 passing. Given this, music which lacks a clear and consistent pulse, music that is broadly ‘unmetred’,
when combined with a low density of events, frequent silences, or long durations, can lead to a subjective
experience of time moving slowly. In some circumstances, the listener’s subjective experience of time may
be so slow as to evoke conceptually the notion of timelessness, leading the listener to ascribe their
experience of the piece as timeless. Such experiences may occur in those compositions which not only lack a
metre but are also without any sense of pulse.

One example of this can be heard in the opening of George Crumb’s Idyll for the Misbegotten (1985). The C4.P30
drum roll that begins the piece quickly fades into a background murmur, the prolonged duration following
the initial strike of the drum—approximately 10 seconds—resulting in a narrowing of attentional focus as
one waits for something to change. Although not a true acoustic silence, the drum roll that pervades this
19
opening is barely audible, resulting in a perception of silence, but one that is shadowed by the faint sounds
of an ominous rumble at the edges of awareness. Consequently, the listener is left in a state of tense
anticipation, the selective focusing of attention on the lengthening duration of the opening as they wait for
what is to come contributing to an experience of temporal suspension (Fraisse 1963: 216). When a change
nally comes, it does so in the form of a short, quick gesture in the ute, arpeggiating up three notes only to
land on a long, sustained tone, leading into a measured vibrato that rst accelerates and then decelerates in
its pulsations. The opening line of the ute ends with a series of quiet and haunting pitch bends, fading into
a prolonged silence marked once again with the nearly inaudible rumbling of the drums. This pattern

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
repeats throughout the opening passage of Crumb’s Idyll (Figure 4.4). Each time the motion a orded by the
ute is eeting, its rhythms failing to evoke any clear sense of pulse or repetition that would a ord
entrainment and, by extension, a means of measuring time.

Figure 4.4

An excerpt from An Idyll for the Misbegotten by George Crumb ©1985 by C. F. Peters. C4.F4

All rights reserved. Used with permission.

In Messiaen’s ‘Louange à l’éternité de Jésus’, evocations of timelessness were created by sustained tones C4.P31
p. 71 that exceeded the bounds of the perceptual present, accompanied by a markedly slow and measured
pulse that itself lent a contemplative air. Within Crumb’s Idyll, timelessness is evoked through extremely
long silences and the absence of any clearly repetitive duration. The sparse texture of this opening,
characterized by short and rapid ourishes leading to long, sustained tones and haunting gestures in the
ute which weave in and out of moments of silence, creates a di erent quality of timelessness, one that is
less contemplative and more primordial, contributing to a sonic landscape that seems to stand outside time.
The use of both ute and drums, instruments Crumb associates with nature (1986), contributes to the
20
evocation of a primitive wilderness that further colours the temporal quality of the movement. The
perceptual experience a orded by this combination of elements is one of time slowed; the semiologically
represented temporal concept one of a fragile moment removed from the ravages of time’s passage.

As I have demonstrated here, invocations of time in music can take on many di erent forms, and the claims C4.P32
made by scholars about musical time can have di erent ontological origins. There is no question that music
is an inherently temporal art form, one comprising sonic phenomena which are brought into existence,
changing, shifting, and then gradually decaying in and through time. Yet it is also possible for music to stop
time, to stand outside of time, to create for the listener an experience of timelessness. Such moments may
come about through the semiological representation of temporal concepts associated with certain musical
features that may be interpreted di erently at di erent points in history, or musical characteristics that tap
into broader psychological principles on what a ects our subjective experience of time passing, allowing the
listener to interpret their experience of music as one of timelessness or stasis. By understanding the
di erent origins of these assertions and the ways in which our interpretation of ongoing perception
in uences our conceptual notions of time, we can arrive at a more ne-tuned understanding of how music
can make time stop.

Notes
1. The fact that sound comes into existence and decays over the course of a duration, and thus depends upon time for both C4.N1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
the perception and processing of sound, is su icient evidence for many that music is temporal and consequently that
questions of timelessness or stasis in music are irrelevant.

2. It should be noted that the concept of timelessness is being treated as an umbrella term that encompasses related C4.N2
concepts such as temporal stasis, an endless present, extended duration, or arrested temporality.

3. The idea that there are multiple di erent types of time can be found in the work of a number of scholars, perhaps the C4.N3
most notable being J. T. Fraserʼs hierarchical theory of time (1978), which is predicated on his belief that time has evolved
over the history of the universe, leading him to specify di erent evolutionary levels of time (1978).

p. 72 4. The assertion that time flows evenly and in equal units is problematic, as it implies the need for another time, a C4.N4
ʻhypertimeʼʼ to measure the flow of the first time in order to assert that it is, in fact, flowing equably. This quickly creates
an infinite and vicious regress as this new ʻhypertimeʼ needs a ʻhyper-hypertimeʼ to measure it, creating a never-ending
cycle of nested times which measure one another (Gunn 1929).

5. Of course, viewing time as measurable is also crucial to science and technology. However, it is worth noting that the unit C4.N5
of time used for measurement is not absolute but relational. E.g. the unit of a second was, at one point, defined as
1/86,400 of the mean solar day. With the development of the atomic clock in the 1960s, the basis for the duration of a
second was changed to the transition between two levels of the ground state of cesium. This definition was then amended
in 1997 to clarify that it applied to the cesium atom in its ground state at a temperature of 0 K. To put it simply, the rate at
which time flows based on the ʻclockʼ is collectively agreed upon and predicated on observable change, not objectively
predetermined.

6. For studies on the role of proportion and the Golden Ratio in music by Béla Bartók and other contemporary composers, C4.N6
see Howat (1977), Kramer (1973), Lendvai (1971), and Lake (2005). For research on the relevance of the Golden Ratio in
tonal music, see Rogers (1981) and Howat (1998). Also of interest is Trevor de Clercqʼs work on the importance of absolute
time, specifically the duration of 2 seconds, for the length of measures in popular music (2016).

7. To the best of my knowledge, while the terms Satz and Gang come from A. B. Marx, it is Hatten who first imbues them with C4.N7
temporal connotations.

8. For Reiner, the poietic dimensions references anything associated with the creation of musical objects that then function C4.N8
as traces (scores, sounds, installations) of musical time, while the esthesic dimension encapsulates a range of responses
to traces of musical time, including the perception, interpretation, or experience of musical time. For a detailed discussion
of his semiotic model of musical time, see ch. 3 of his book, Semiotics of Musical Time (2000).

9. Within Critique of Pure Reason, Kant argues: ʻTime is not something that would subsist for itself or attach to things as an C4.N9
objective determination […] time is therefore merely a subjective condition of our (human) intuition […] and in itself,
outside the subject, is nothingʼ (1781(1998): 163–164).

10. For a discussion of factors that may contribute to di erences between perceived and notated tempo, see London (2012). C4.N10

11. One need only read Thomas Mannʼs Magic Mountain (1824) (or any number of other novels, for that matter) or observe the C4.N11
various manipulations of time in film that are then readily accepted by the audience to recognize that the concepts of
timelessness, stasis, and a suspended present are relevant to the Westernized individual.

12. The role and experience of performers is another perspective that can be considered within the framework presented C4.N12
here. Prior research has shown that many performers experience flow states, which are themselves associated with losing
track of time and a sense of a suspended ongoing moment that has no boundaries (Csíkszentmihályi 1990). As performers
would arguably embody a di erent kind of relationship to music and its temporalities, I have chosen to focus exclusively
on listeners within the context of this chapter.
13. At this point in time, it is unknown what factors play a role in one individual being more susceptible to experiences of C4.N13
timelessness over another. Very likely personality aspects, such as absorption in music—which are known to play a role in
p. 73 the induction of emotion while listening to music (Hogue et al. 2016) as well as the perceived emotionality of the music
itself (Noulhiane et al. 2007)—may be factors in the induction of timelessness while listening to music.

14. The actual duration of the perceptual present is variable, and depends upon factors such as attention, the number and C4.N14
complexity of the events one is experiencing, and the degree to which those events cohere into gestalts and form
hierarchical relationships (Fraisse 1963; Zakay 1989). However, given the sparseness of Messiaenʼs melodic line and the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
slow rate of change, it is more likely that the duration of the perceptual present for someone listening to this work sits in
the lower bounds of the possible range.

15. The amount of vibrato used by the performer will influence the perceptual e ect of this opening, as a high amount of C4.N15
vibrato will provide a temporal movement within the sustained tones that may become a focus for the listener,
subsequently downplaying the experience of extreme slowness a orded by the cello line.

16. For a related analysis that focuses on the temporal e ect of the pulsed piano chords, see Noble (2017). C4.N16

17. A 2005 study by Wearden demonstrates this phenomenon by assigning participants to either spend 9 minutes watching C4.N17
the movie Armageddon or sitting in a simulated waiting room. Unsurprisingly, participants in the Armageddon condition
reported experiencing time moving faster than normal, while those in the waiting room reported experiencing time
dragging. However, when a second group of participants were asked to provide temporal judgements on their condition
a er spending time on an intermediary task, the results were reversed: those in the Armageddon condition judged the
duration longer than did those in the waiting room. Such a finding underscores di erences between the experience of
time in the moment versus recollecting the temporal quality of an experience from memory.

18. An empirical study exploring the influence of silences in contemporary music on the experience of time found that C4.N18
silences resulted in a perception of time moving more slowly than did motives (Knowles and Ashley 2017).

19. Elizabeth Margulis (2007) di erentiates between acoustic and perceived silences, noting that in certain circumstances C4.N19
there may be sound on an acoustic level that nevertheless results in a perception of silence for the listener.

20. For a more in-depth discussion concerning the representation of time in this work and its relationship to 20th-c. C4.N20
depictions of the pastoral, see Knowles (2017).
References C4.S4

Barry, B. R. (1990). Musical time: The sense of order. Pendragon Press. C4.P33
Google Scholar Google Preview WorldCat COPAC

Bellman, J. D. (2010). Chopinʼs Polish Ballade: Op. 38 as narrative of national martyrdom. Oxford University Press. C4.P34

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
Berger, K. (2007). Bachʼs cycle and Mozartʼs arrow: An essay on the origins of musical modernity. University of California Press. C4.P35
Google Scholar Google Preview WorldCat COPAC

Block, R. A. (1985). Contextual coding in memory: Studies of remembered duration. In J. A. Michon and J. L. Jackson (eds), Time, C4.P36
mind, and behavior, 169–178. Springer.
Google Scholar Google Preview WorldCat COPAC

Block, R. A. (1989). Experiencing and remembering time: A ordance, context and cognition. In I. Levin and D. Zakay (eds), Time C4.P37
and human cognition: A life-span perspective, 333–363. North-Holland.
Google Scholar Google Preview WorldCat COPAC

p. 74 Block, R. A., Hancock, P. A., and Zakay, D. (2010). How cognitive load a ects duration judgments: A meta-analytic review. Acta C4.P38
Psychologica 134(3): 330–343.
Google Scholar WorldCat

Block, R. A., and Reed, M. A. (1978). Remembered duration: Evidence for a contextual-change hypothesis. Journal of C4.P39
Experimental Psychology: Human Learning and Memory 4(6): 656–665.
Google Scholar WorldCat

Block, R. A., and Zakay, D. (1997). Prospective and retrospective duration judgments: A meta-analytic review. Psychonomic C4.P40
Bulletin & Review 4(2): 184–197.
Google Scholar WorldCat

Boltz, M. (1989). Time judgments of musical endings: E ects of expectancies on the ʻfilled interval e ectʼ. Perception & C4.P41
Psychophysics 46(5): 409–418.
Google Scholar WorldCat

Clayton, M. (2000). Time in Indian music: Rhythm, metre, and form in north Indian rāg performance. Oxford University Press. C4.P42
Google Scholar Google Preview WorldCat COPAC

Cli on, T. (1983). Music as heard: A study in applied phenomenology. Yale University Press. C4.P43
Google Scholar Google Preview WorldCat COPAC

Crumb, G. (1986). An idyll for the misbegotten: Amplified flute and percussion, 3 players/George Crumb. C. F. Peters. C4.P44
Google Scholar Google Preview WorldCat COPAC

Csikszentmihalyi, M. (1990). Flow: The psychology of optimal experience. Harper & Row. C4.P45
Google Scholar Google Preview WorldCat COPAC

Delaere, M. (2009). Tempo, metre, rhythm, time in twentieth-century music. In D. M. Crispin (ed.), Unfolding time: Studies in C4.P46
temporality in twentieth-century music, 13–43. Leuven University Press.
Google Scholar Google Preview WorldCat COPAC

Epstein, D. (1981). On musical continuity. In J. T. Fraser, N. Lawrence, and D. Park (eds), The study of time IV: Papers from the C4.P47
Fourth Conference of the International Society for the Study of Time, 180–197. Springer.
Google Scholar Google Preview WorldCat COPAC
Fraisse, P. (1963). The psychology of time (trans. J. Leith). Eyre & Spottiswoode. C4.P48
Google Scholar Google Preview WorldCat COPAC

Fraser, J. T. (1978). Time as conflict. Birkhäuser. C4.P49


Google Scholar Google Preview WorldCat COPAC

Frith, S. (1996). Performing rites: On the value of popular music. Harvard University Press. C4.P50
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
Gri iths, P. (1985). Olivier Messiaen and the music of time. Faber. C4.P51
Google Scholar Google Preview WorldCat COPAC

Gunn, J. A. (1929). The problem of time. Allen and Unwin. C4.P52


Google Scholar Google Preview WorldCat COPAC

Hasty, C. (1997). Meter as rhythm. Oxford University Press. C4.P53


Google Scholar Google Preview WorldCat COPAC

Hatten, R. (1997). Music and tense. In Semiotics around the world: Synthesis in diversity, 627–630. Mouton de Gruyter. C4.P54
Google Scholar Google Preview WorldCat COPAC

Hogan, H. W. (1975). Time perception and stimulus preference as a function of stimulus complexity. Journal of Personality and C4.P55
Social Psychology 31(1): 32–35.
Google Scholar WorldCat

Hogue, J. D., Crimmins, A. M., and Kahn, J. H. (2016). ʻSo sad and slow, so why canʼt I turn o the radioʼ: The e ects of gender, C4.P56
depression, and absorption on liking music that induces sadness and music that induces happiness. Psychology of Music 44(4):
816–829.
Google Scholar WorldCat

Howat, R. (1977). Debussy, Ravel and Bartok: Towards some new concepts of form. Music & Letters 58(3): 285–293. C4.P57
Google Scholar WorldCat

Howat, R. (1998). Architecture as drama in late Schubert. In Schubert Studies, 166–190. Ashgate. C4.P58
Google Scholar Google Preview WorldCat COPAC

Hyland, A. (2016). In search of liberated time, or Schubertʼs Quartet in G major, D. 887: Once more between sonata and variation. C4.P59
Music Theory Spectrum 38: 85–108.
Google Scholar WorldCat

Ihle, R. C., and Wilsoncro , W. E. (1983). The filled-duration illusion: Limits of duration of interval and auditory fillers. Perceptual C4.P60
and Motor Skills 56(2): 655–660.
Google Scholar WorldCat

Jones, M. R., and Boltz, M. (1989). Dynamic attending and responses to time. Psychological Review 96(3): 459–491. C4.P61
Google Scholar WorldCat

Klein, M. (2004). Chopinʼs fourth ballade as musical narrative. Music Theory Spectrum 26(1): 23–56. C4.P62
Google Scholar WorldCat

Knowles, K. (2017). A broken Idyll: Post-pastoralism in the works of George Crumb. E-Rea 14(2): 1–15. C4.P63

p. 75 Knowles, K., and Ashley, R. (2017). Time versus tension: Studying the influence of contemporary music on experienced time. C4.P64
Conference presentation, SMPC 2017.
Google Scholar Google Preview WorldCat COPAC

Kramer, J. D. (1973). The Fibonacci series in twentieth-century music. Journal of Music Theory 17(1): 110–148. C4.P65
Google Scholar WorldCat

Kramer, J. D. (1988). The time of music: New meanings, new temporalities, new listening strategies. Shirmer Books. C4.P66
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
Lake, W. E. (2005). Form and temporal proportions in George Crumbʼs ʻAncient voices of childrenʼ. In S. Bruns, O. Ben-Amots, and C4.P67
M. D. Grace (eds), George Crumb & the alchemy of sound: Essays on his music, 83–100. Colorado College Music Press.
Google Scholar Google Preview WorldCat COPAC

Langer, S. (1953). Feeling and form; a theory of art. Scribner. C4.P68


Google Scholar Google Preview WorldCat COPAC

Lendvai, E. (1971). Béla Bartók: An analysis of his music (ed. A. Bush). Kahn & Averill. C4.P69
Google Scholar Google Preview WorldCat COPAC

Leong, D. (2001). A theory of time-spaces for the analysis of twentieth-century music: Applications to the music of Béla Bartók. C4.P70
University Microfilms International.
Google Scholar Google Preview WorldCat COPAC

Leydon, R. (2003). ʻCes nymphes, je les veux perpétuerʼ: The post-war pastoral in space-age bachelor-pad music. Popular Music C4.P71
22(2): 159–172.
Google Scholar WorldCat

London, J. (2012). Hearing in time: psychological aspects of musical meter (2nd edn). Oxford University Press. C4.P72
Google Scholar Google Preview WorldCat COPAC

Luchese, D. (2010). Olivier Messiaenʼs slow music: A reflection of eternity in time. In J. Crispin and L. Sitsky (eds), Collected Work: C4.P73
Olivier Messiaen, 179–192. Cambridge Scholars.
Google Scholar Google Preview WorldCat COPAC

Margulis, E. H. (2007). Moved by nothing: Listening to musical silence. Journal of Music Theory 51(2): 245–276. C4.P74
Google Scholar WorldCat

Margulis, E. H. (2014). On repeat: How music plays the mind. Oxford University Press. C4.P75
Google Scholar Google Preview WorldCat COPAC

Marinelli, P. (1971). Pastoral. Methuen. C4.P76


Google Scholar Google Preview WorldCat COPAC

McClary, S. (2012). Desire and pleasure in seventeenth-century music. University of California Press. C4.P77
Google Scholar Google Preview WorldCat COPAC

Monelle, R. (2000). The sense of music. Princeton University Press. C4.P78


Google Scholar Google Preview WorldCat COPAC

Noble, J. (2017). What can the temporal structure of auditory perception tell us about musical ʻtimelessnessʼ? MS submitted for C4.P79
publication.
Google Scholar Google Preview WorldCat COPAC

Noulhiane, M., Mella, N., Samson, S., Ragot, R., and Pouthas, V. (2007). How emotional auditory stimuli modulate time C4.P80
perception. Emotion 7(4): 697–704.
Google Scholar WorldCat
Ornstein, R. (1969). On the experience of time. Penguin Books. C4.P81
Google Scholar Google Preview WorldCat COPAC

Peattie, T. (2002). In search of lost time: Memory and Mahlerʼs broken pastoral. In Collected work: Mahler and his world, 185–198. C4.P82
Princeton University Press.
Google Scholar Google Preview WorldCat COPAC

Reiner, T. (2000). Semiotics of musical time. Peter Lang. C4.P83

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469071 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Rogers, M. R. (1981). Chopin, Prelude in A minor, Op. 28, No. 2. 19th-Century Music 4(3): 245–250. C4.P84
Google Scholar WorldCat

Short, R. (2019). ʻA-five, six, seven, eight!ʼ Musical counting and dance hemiolas in musical theatre tap dance breaks. Conference C4.P85
presentation, SMT 2019.
Google Scholar Google Preview WorldCat COPAC

Taylor, B. (2016). The melody of time: music and temporality in the Romantic era. Oxford University Press. C4.P86
Google Scholar Google Preview WorldCat COPAC

Thurmaier, D. P. (2006). Time and compositional process in Charles Ivesʼs Holidays Symphony, Doctoral dissertation, Indiana C4.P87
University.

Zakay, D. (1989). Subjective time and attentional resource allocation: An integrated model of time estimation. In I. Levin and C4.P88
p. 76 D. Zakay (eds), Time and human cognition: A life-span perspective, 365–397. North-Holland.
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
CHAPTER

5 Rhythm, Time, and Presence  C5

Anne Danielsen

https://doi.org/10.1093/oxfordhb/9780190947279.013.8 Pages 77–C5.P81


Published: 08 December 2021

Abstract
Repetitive rhythm-oriented or groove-based music is speci cally designed to usher the listener into a
state of absorption or presence in the music. This chapter argues that this particular form of musical
time originates with the repetition of a certain inner dynamics within the basic rhythmic pattern of the
groove. The rst part of the chapter presents structural and microrhythmic features that seem to be
particularly impactful upon the listener’s state of groove absorption. The second part discusses the
consequences of the fact that it is impossible to describe such an experience of musical time without
stepping outside it. An initial observation is that in the very moment one starts to consider the state of
being-in-the-groove, one’s state of being is changed.

Keywords: rhythm, timing, groove, performance, presence, repetition, absorbed state of being
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

The gift of presence is the property of the event. C5.P1

Martin Heidegger C5.P2

RHYTHM is a fundamentally temporal phenomenon. It has been de ned as ‘an ordered alternation of C5.P3
contrasting elements’ (Crossley-Holland 2017), ‘a regularly recurring sequence of events or processes’
(‘Rhythm’, n.d.), or as ‘articulated ow’ (Hasty 1997: 68). In a musical context, it is often understood in
relation to metre, i.e. as a regular, repeated pattern structured according to a certain matrix of strong and
1
weak beats. Common to all these understandings of rhythm is the fact that, when experiencing rhythm, the
listener applies a ‘structure’ to the unfolding of time. It is this structuring capacity in the listener that is
2
necessary for the transformation of everyday time to musical time and its rhythms.

In what follows, I present some basic insights of research into musical rhythm, and then give examples of C5.P4
structural and microrhythmic features that seem to be particularly successful in bringing the listener into
the state of being-in-the-groove—that is, the particular presence in the now that is often associated with
groove experiences. By ‘microrhythm’, I here refer to the overall rhythmic shaping of musical events at the
micro level, encompassing timing, as well as duration, shape, timbre, and intensity. In the second part of
the chapter, I will delve further into the relationships among rhythm, time, and presence, addressing the
role of repetition and the inner dynamics of groove patterns with regard to the listener’s state of groove
absorption. Finally, I will undertake an epistemological re ection upon the challenges of capturing the
p. 78 groove experience as it actually unfolds in time. A central topic pertaining to this re ection, as well as to
the twofold approach of this chapter overall, is the tension between the study of rhythm as an analytical
object, having objecti ed cultural or musical meaning, which is bound to take place in language and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
‘outside’ of the time of experiencing it, and the study of rhythm from ‘inside’ of the not-as-yet
communicated experience, i.e. as a presence in time.

Rhythm as Process C5.S1

Most researchers agree that musical rhythm as an experiential phenomenon comes into being in the C5.P5
meeting of sound and listener (see e.g. Bengtsson et al. 1969; Clarke 1985; 1987; 1989; Danielsen 2006; 2010;
Desain and Honing 1989; Kvifte 2004; London 2012)—that is, rhythm happens when listeners hear the ow
of sound in time as rhythmic. This process depends on both the features of the sonic ‘stream’ and a
structuring capacity in the listener: when the features of the stream are such that they enable it to be heard
as rhythmic, schemes are generated or activated in the perceptual and cognitive apparatus of the listener.
Such schemes rely on basic oscillating mechanisms in the perceiver that entrain to the pulse of the music
(Large and Jones 1999; Jones 1976), as well as learned patterns related to more style-speci c aspects of the
music. The latter is part of the listener’s pre-understanding—that is, what the listener brings to the act of
listening through training and culture. The schemes, or reference structures (Danielsen 2010), are used by
the listener to make sense of the organization of the music and to develop expectations for its course. They
are virtual, in the sense that they do not exist as sound per se: they are activated in the listener by the actual
sound that unfolds in time, to be used in the act of structuring the same sound. For example, we perceive a
tempo (how fast is the pulse of the music?) and metre (what is the underlying recurring virtual matrix of
durations and accentuations?), and recognize the basic rhythmic gures (what pattern(s) is/are being
played?) as well as the style and genre.

The interaction of the actual and virtual aspects of rhythm is often conceived of as an interaction of C5.P6
sounding rhythm and the hierarchical organization of levels of metre as depicted in traditional Western
notation (see e.g. London 2012), However, if we want to transcend the framework of traditional music
theory, we might say that the experience of rhythm happens in the meeting of sound and a multidimensional
framework of reference structures that might comprise everything from metre to rhythmic gures, patterns of
timing and phrasing, bodily gestures—in fact, any conceivable temporal mechanism used in the structuring
process. Accordingly, the speci c structures that are induced in the listener derive from a combination of
general perceptual processes, the listener’s pre-understanding—i.e. what the listener brings to the act of
listening through training, culture, and so on—and the schemes that are implicitly ‘present’ in the actual
sounding rhythmic events. In principle, everything that is used to structure the course of the music is a
reference structure. Bodily gestures, for example, which are often regarded as externalizations of such
p. 79 reference structures, are probably also constitutive of them, as they represent an embodied means of
transforming sound from an undi erentiated sonic stream in time to an understandable rhythm (Danielsen
2010; Danielsen et al. 2015; Haugen 2016).

Reference structures generate expectations or predictions in the listener regarding the events to come, and C5.P7
a crucial aesthetic aspect of rhythm, then, is the playing with, or transformation of, such expectations over
time. We might get what we expect, or we might not. In other words, sounding rhythm contains both its
relevant structuring pattern and the potential for a signi cant or expressive variation of this pattern.
Maintaining the correct balance between expected and unexpected events and variations upon such events
is crucial in groove-based music, and also (as we shall see) key to keeping the listener engaged in the
groove. In the following, I focus on some structural and microrhythmic features that are particularly
striking with regard to drawing the listener into the groove.

The rst feature is the structural tension between the main rhythm and some competing layer of pulses, C5.P8
such as articulating four strokes against only three beats of the pulse. In James Brown’s classic tracks from
the funk era, such 4:3 patterns can be played by the guitar, the bass, the bass drum, the horns, or a

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
combination of several instruments (see Danielsen 2006: 61–72). It is an aspect of songs like ‘Sex Machine’,
‘Soul Power’, ‘Superbad’, and ‘The Payback’, to mention a few (for analyses, see Danielsen 2006; Wilson
1974). In ‘Soul Power’, for example, we see that an incomplete version of the 4:3 gure is present in bars 1
and 2 of the repeated basic unit of the groove. It starts on the second beat and ends by opening out toward
the fourth (Figure 5.1).

Figure 5.1

Cross-rhythmic counter-rhythm in ʻSoul Powerʼ. C5.F1

The many variants of this gure work as a more or less explicit counter-rhythm to the basic pulse. I call C5.P9
them ‘cross-rhythmic’ counter-rhythms because they derive from a competing pulse that, if it were to be
continued throughout the groove, would generate a true cross-rhythm (Nketia 1974: 134–135)—i.e. a
rhythmic texture featuring two equally important basic pulses at the same time. However, in most of the
popular music styles in which this feature is common—other examples include EDM and Afro-Cuban dance
music—this competing pulse is never allowed to take over the main metre, so there is only a ‘tendency of
cross-rhythm’ (Danielsen 2006: 62–66). In ‘Soul Power’, it is particularly striking that the gure only hints
at a competing pulse and does not set up a separate layer of counter-rhythm or competition with the main
beat. Instead, it acts to both animate and break up the primary metre.

Another ‘class’ of counter-rhythms are ‘on-and-o ’ gures, which arise, for example, when strokes are C5.P10
p. 80 rst played on and then o the beats (see Figure 5.2). This gure can be found in the famous bridge from
James Brown’s ‘Sex Machine’ (see Danielsen 2006: 80–82).

Figure 5.2

On-and-o figure in the bridge of ʻSex Machineʼ. C5.F2

Often these counter-rhythmic patterns, the 4:3 and the on-and-o , are placed so that they destabilize the C5.P11
groove at the end of the groove pattern—that is, when the metrical structure leads the pattern toward a
‘mini-closure’. Inducing such instability at the natural point of metric closure has a driving, dynamic e ect
and propels the process onward.
Both are also examples of a form of structural tension within the groove that generates double reference C5.P12
structures in the listener. In the case of the 4:3 pattern, two di erent pulse schemes with tempi in the ratio
4:3 arise; in the case of the on-and-o pattern, the result is two pulses in the same tempo but in an anti-
phase relationship.

Ambivalence and tension at the micro level of rhythm may also keep the listener engaged in the groove. C5.P13
Parliament’s live recording of ‘P-Funk (Wants to Get Funked Up)’ (P-funk All-Stars Live at the Beverly

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
Theater in Hollywood, Westbound Records, 1990) is one example. In the two bar ‘P-Funk’ ri from this
recording, we nd two versions of the same rhythmic gure (A and B in Figure 5.3). The rst (A) starts at
beat 2 and ends at the second sixteenth note of beat 3. The ri continues in bar 2 with a second version of
the gure (B) now starting from beat 1 and ending on the second sixteenth note of beat 2 (Figure 5.3a).

Figure 5.3a

Schematic outline of the first phrase of the melody-ri of ʻP-Funkʼ (live version). C5.F3

Looking at a waveform representation of the performed ri (Figure 5.3b), we see that several of the C5.P14
sixteenth-note syncopations are performed slightly swung. The fourth note of version A of the ri , for
example, instead of hitting at the second sixteenth note (3i), falls right before the straight 3& position (see
Figure 5.3b). The accurate location of these syncopations are, however, ambiguous because the accurate
positions of the basic 4/4 beats are rather open at the micro level: In bar 1, the I-O-I between beats 1 and 2
(beat 1.1 to beat 1.2) is considerably shorter than the corresponding virtual beats in bar 2 (2.1 to 2.2). This
p. 81 means that the fourth stroke of the ri falls late in relation to the pulse established at the beginning of
bar 1, but early when repeated after the slightly slower pulse established by the I-O-I from beat 1.4 to beat
2.1—even though it is actually earlier if measured against the isochronous grid indicated by the black dotted
lines in Figure 5.3b. The openness as to the correct location of such strokes is also enhanced by the
audience’s handclapping on beats 2 and 4, which smears out the pulse reference, transforming it to beat bins
(Danielsen 2010; 2018) of about 100 milliseconds.
Figure 5.3b

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
Waveform (top) and spectrogram (bottom, 0–12000 Hz) of the first phrase of the melody-ri of ʻP-Funkʼ (live version). (Produced C5.F4
in Praat, v. 6.0.46, praat.org). Isochronous grid of quarter notes indicated by vertical, black dotted lines. To view the figure in

colour, please visit the companion website

Summing up, the appeal of the ri of ‘P-Funk (Wants to Get Funked Up)’ probably lies in a combination of C5.P15
(a) the structural phase-shift of the gure from bar 1 to bar 2 (the gure is unaltered, but starts from
di erent positions in the bar), (b) the ambiguity produced through the many small-scale timing
adjustments in the performance of the rhythmic pattern, and (c) the general openness as regards metric
positions at the micro level in this groove. The latter is related to the loosely located ‘cluster-beat’ pulse (i.e.
the handclaps of the audience), which establishes a range of possible pulse locations. However, the example
also illuminates the ways in which timing in ections by musicians in performance might work to keep us
engaged in the groove. Although empirical results from research into microtiming are mixed (see Davies et
al. 2013; Matsushita and Nomura 2016; Senn et al. 2016), such timing and other microrhythmic features are
p. 82 thought by musicians, musicologists, and ethnomusicologists alike to be very responsible for the drive of
a groove (see e.g. Butter eld 2006; Iyer 2002; Keil 1994a; 1994b; 1995; Kilchenmann and Senn 2015;
Monson 1996; Prögler 1995).

All these analytical examples demonstrate the very dynamic nature of rhythm. Several reference- C5.P16
structuring principles can be at work at the same time, and not only the sounds themselves but also the
structures used to understand them can be subject to continuous change. Moreover, metrical and non-
metrical reference structures that relate to basic perceptual processes, cultural patterns, and the particular
sounds presented to the listener will be actualized in the listener by the music, perhaps changed as a
consequence of the listening experience, and furthermore lodged in the listener thereafter, to be readily
available for new experiences. How this process unfolds—i.e. what reference(s) we use to structure its
events at each point in time—will decide what groove we actually hear.

In a groove, interestingly, the same structural or microrhythmic ambivalence is repeated each time the C5.P17
basic unit is repeated. There is, in other words, a stable instability present that is aesthetically satisfying.
This becomes even more conspicuous when one approaches looped computer-based grooves—that is,
grooves where each basic unit, each repetition, is exactly identical to the previous one. In these grooves, as
well, it is important to compose the basic unit in such a way that a compelling inner dynamic drives the
groove forward. And, as with the performed grooves discussed above, this can be done either by combining
di erent layers of rhythm at a structural level or by introducing microrhythmic tension (see e.g. analysis in
Danielsen 2015).

In all these examples, then, it is not the di erence between one repetition and the next that matters, but C5.P18
rather the productive dissymmetry that works from inside the repeated pattern and is experienced from
inside the process. In music that is being performed, this repetitive production is re ected in the fact that a
musician has to produce every beat, and, from this perspective, every beat is actually new. In computer-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
based grooves, this is not the case, since the exact same sound is often used in every bar. Yet time is still new.
When one is absorbed in the now of the experience, one will move together with the groove from one
gesture to the next. This means that, while every new repetition no longer has to be performed by
musicians, sounds still need to be perceived, time and again. One could even go so far as to say that a
presupposition of the state of ‘being-in-the-groove’ is that repetition is not heard—that it is out of focus.
From a phenomenological perspective, every repetition should be brand new.

A second insight to be drawn from the above analyses is that it is not repetition in itself that draws the C5.P19
listener into the state of being in a groove. A metronome, for example, rarely has this e ect. Repetition in a
groove context means, instead, repetition of the inner dynamics of the groove’s basic unit, and this inner
dynamic is often designed to keep the listener engaged throughout the entire pattern. The structural
tensions inherent in groove patterns of the James Brown and P-Funk grooves presented above contribute to
a productive dissymmetry, and its certain driving, dynamic e ect leads the process onward. If we take the
gure of on-and-o as an example, the listener is almost forced to follow the pattern from on to o and
back again. From the rst confusion surrounding the critical point where the gure is inverted, through the
p. 83 part of the pattern where the gure actually is inverted, to the point where everything is back in place
again, the listener is kept busy at the perceptual level with sorting out the interplay of rhythm and counter-
rhythm. When this phase of relative stability arrives in the latter part of the pattern, as is the case with ‘Sex
Machine’, the pattern contains a certain built-in circularity. The metrical structure of the main rhythm that
leads the pattern toward a ‘mini-closure’ at the end of every basic unit is neutralized by the counter-
rhythm. This is typically also the case with the cross-rhythmic gure: it increases the
dissymmetry/instability towards the end of the basic unit and is directly or indirectly at work when the
natural point of closure arrives, and perhaps also slightly after it (it usually takes some time to get things
back in order).

Rhythm as Presence C5.S2

We have now discerned some typical features of rhythm and discussed the ways in which rhythms tend to C5.P20
keep the listener invested in the process, noting along the way that it is the repetition of an inner dynamic of
structural and microrhythmic ambivalences, such as those described above, that is the primary force behind
the state of being-in-the-groove. None of the above, however, is an attempt to describe this state of being
as such—that is, how it is experienced when one is there, present in the rhythmos, the peculiar ow of time
that rhythm is. Analytical investigations are bound to take place at a distance; they require a presence in the
analytical now, whereas a description of what rhythm actually is when present in the now of the groove is
something di erent (albeit related). The latter, which is a real phenomenological account of the state of
being-in-the-groove, implies the study of rhythm not as an analytical object or objecti ed cultural or
musical meaning but as lived experience—i.e. as an experience prior to (as well as following, of course) our
speech about it.

There are, however, major challenges linked to such an ambition, since in principle it transcends what C5.P21
might be called the ‘domain of thought and meaning’. In the following, I re ect on the dialectics of event
and meaning and the ways in which any experience is fundamentally conditioned by the temporal unfolding
of these two aspects of the process of understanding being.

A central issue in phenomenology and contemporary hermeneutics is the temporal character of human life, C5.P22
and especially the temporal space between immediate ‘being in time’ and the re ective understanding of
3
this same being. One result of the fundamentally temporal character of being is that the exchange between
being and understanding, two basic modes of experience, is in itself bound to unfold in time. Because these

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
modes cannot take place simultaneously—there is no way of attending to one’s own experience at a
distance—there is no possible way of arriving at a systematic understanding of being as such. Being or
subjective experience is unavoidably objecti ed through the process of distanciation. Consequently, every
attempt to understand being implies transforming what is to be understood into something other than what
p. 84 it was. Even the very act of framing/naming rhythm as a state of being, then, tends to obscure the
distinctive qualities of its experiential form.

This process of objecti cation concerns the problem of groove—or any other event, for that matter— C5.P23
coming to one’s senses. Individual, subjective experience is in fact the only access we have, not only to the
subjective aspects of musical experience but to its intersubjective aspects as well—we are simply left to
understand, and communicate, our experiences after the fact. The di erence between how things are when
they happen, in time, and how they are immediately afterward, in the process of understanding, is
impossible to transcend: it cannot be levelled out. At the same time, there are di erences among various
discourses regarding the extent to which they downplay, or conversely attempt to highlight, their
underlying bases in the temporal unfolding of subjective being.

The Imploding Now: Feeling and Intensity C5.S3

In his lecture ‘The hermeneutical function of distanciation’, Ricœur proposes an interpretation of discourse C5.P24
as a dialectic between performing and understanding, or, in his words, event and meaning. According to
Ricœur, to say that discourse is an event is to say several things at the same time. Discourse is realized
temporally and in the present: it ‘has time’, as Ricœur says. Furthermore, it is characterized by the fact that
someone is speaking: it has subjectivity. Finally, the speaking takes place and refers to a world: it includes
others (Ricœur 1973: 130–131). The other pole of the dialectic—meaning—is equally important to the
constitution of discourse. And in line with the hermeneutic tradition beginning with Heidegger, Ricœur
situates distanciation within being-in-the world, or, as he writes: ‘The moment of understanding responds
dialectically to being-in-a-situation’ (Ricœur 1973: 140).

Ricœur’s notion of meaning is tied to the unavoidable distanciation that is a part of discourse itself, and the C5.P25
4
very rst distanciation is, as Ricœur writes, ‘the distanciation of the saying in the said’ (Ricœur 1973: 132).
Ricœur here opens up the notion of meaning way beyond the more common use of meaning as Bedeutung,
implying all of the aspects and levels of exteriorization that make the inscription of an event possible. By
this he emphasizes that inscription is unavoidable, but may take place on di erent levels.

With the speech act theory of Austin and Searle as a starting point, Ricœur analyses discourse as potentially C5.P26
activating three levels of meaning. The rst level is the locutionary act, or the propositional meaning, in the
sense of ‘what is said’. The second level is the illocutionary act or force, ‘that which we do in saying’, or
rather, how something is said. The third level, the perlocutionary act, is located in and with the very act of
uttering; it is the fact that something is being said. Even in those cases where meaning is present solely at
p. 85 the level of perlocutionary action, even though the meaning is no more than a ‘that something was’, the
event does not escape distanciation. The event will always be turned into some form of inscription.
Following Ricœur, one might say that the state of being-in-the-groove is also actualized as event and C5.P27
understood as meaning. There is always a dialectical response to the being-in-the-situation. The
paradoxical relationship of the oating temporal experience and its identi able meaning, referred to by
Ricœur as the dialectic of event and meaning, is a fundamental condition of being in general. In other
words, what Ricœur points out regarding discourse might also be pointed out about music: all music is
actualized as event; all music is understood as meaning. The relationship among the three levels of meaning
may, however, di er considerably from the meaning of discourse. The hierarchy should probably be turned

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
upside down: due to music’s lack of capacity to produce a referential meaning in a linguistic sense, the
locutionary act (what is said) is regarded as relatively unimportant. The illocutionary level (how we say) is,
however, very important. Ultimately, though, the perlocutionary action (the act of saying)—the way music
‘can stir up certain e ects’, to paraphrase Ricœur, and have a ‘direct in uence upon the emotions and the
a ective dispositions’—is probably the bottom line in many cases.

What, then, does the immediate perlocutionary action amount to in the case of groove? What is groove as it C5.P28
happens? Using the dialectic of event and meaning as a starting point, one might say that the state of being-
in-the-groove is characterized by an intensi cation of the event as event—the state of meaning is never
allowed to arrive; instead, it is postponed or set aside. It is put out of play by the play, so to speak, of the
music. Instead of responding dialectically to being-in-the-situation with a moment of understanding, the
state of being-in-the-groove means that we nd ourselves absorbed in a delay of meaning. The presence in
the event defers the presence of meaning. This intensi cation can be understood as a prolonged presence in
the event. Instead of stepping out of time to sum up the course of the music, you remain in the course,
moving together with it through time. The ways in which music is structured play a role in whether the
listener will remain in the course of the events (i.e. in time) or step out of time to form a larger structure. A
classical cadence, for example, can be thought of as a point in time where such an overview of a larger chunk
of the musical course, i.e. musical form, is created.

Apart from such moments of intensity, however, one is completely engaged in the course of groove as such. C5.P29
Gilles Deleuze’s distinction between dynamic and static repetition (Deleuze 1994), which corresponds
respectively to repetition as experienced from a position inside and outside the process, may shed light on
this. When experiencing the groove from within the process, a repetition—the pattern or gure that is
repeated—is not shaped at once but rather comes into being like the ‘evolution’ of a bodily movement. In
parallel, when in the state of being-in-a groove, the rhythmic pattern is not experienced at once but instead
over time. Every part of the rhythm has to be e ected each time: one makes up one part and answers with
another in an eternal rhythmic dialogue.

It is this repetition of an internal dynamics or di erence within that which is repeated, that Deleuze C5.P30
p. 86 describes as ‘productive dissymmetry’. The examples presented here, such as the gures of four-
against-three or on-and-o beats, contribute to such a dissymmetry at the structural level since it requires
a constant structuring and restructuring of the ongoing events. Timing in ections might also be thought to
work in this way, either by providing variations from one basic unit to the next or by so-called systematic
variation (Bengtsson et al. 1969), i.e. a productive dissymmetry within each unit.

The intensi cation of the event character of groove may thus be conceived of as a reversal of the delay C5.P31
5
caused by the circulation of signs. The intensi cation is a delay in the opposite direction: what is deferred is
the very entrance into the circulation of signs. This ‘implosion’ in the event defers the transition to the
moment of understanding, where meaning is allowed to sum up the happenings, for example, by naming
the event ‘an event’. The shift to meaning is completely out of sight. One is in the event as if there were
never anything else—perhaps to such an extent that, even after the fact, one is likely to deny that a shift to
meaning ever took place.
What is—or rather, how is—this particular imploding event in terms of aesthetic experience? It is probably C5.P32
no coincidence that the few words in James Brown’s funk that have a meaning—‘I Feel Good’, ‘Cold Sweat’,
‘I Got the Feeling’, ‘I Am Superbad’, ‘Get Up’, ‘(I Feel Like Being a) Sex Machine’, ‘Get on the Good Foot’,
and so on—deal with groove-based music’s own aesthetic standard: the feel. Feel has nothing to do with an
extrovert style of performance or with expressing actual innermost feelings. It is about cultivating the right
swaying motion, about accurate timing, about being in place, in time, about precision and relaxation at
once, about playing all the correct gures in the right way. However, even though the right feel is present in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
the event, it may remain more or less concealed on the level of meaning. And one may ask: how do I know
that this feeling of feel took place? How, or perhaps rather when, is it that I feel that I feel? If the state of
being in the groove is that of a total absorption, and there is never any distance, one feels, pure and simple.
There is no aesthetic re ection: one does not feel that one feels.

This may be exactly the time for an intervention to take place. James Brown, for example, tends to enter this C5.P33
‘unawareness’, giving the absorbed participant a little shake by bringing us out of the absorption. Musically,
this is often achieved through some form of break, played by (for example) the drums or the horn section. A
break is, like the classical cadence, a salient point in the course of a groove, and by interrupting the musical
movement, it transfers one’s attention to what is going on. When the groove continues with more of the
same after the break, it is nevertheless di erent. The ‘new’ beginning underlines the qualities of that which
was and now is: suddenly, there is a di erence; for a moment, I feel that I feel. However, this di erence does
not imply that this feel of feel necessarily has to be transformed into something else. Instead, it is an instant
of intensi ed feeling, as if we suddenly sense the state of being that we have moved away from, and now
return to, in a new way. Rather than a stop or breakdown, the break is experienced as a heightened sensation
—as intensi ed feeling, or simply intensity.

It is important to note here that this form of intensity is not an increase of intensity, understood as the C5.P34
di erence between the state-of-being prior to the break and after. It is instead an intensifying of one state
of being as such (Deleuze 1994: 232). Before the break, there is only absorption in the groove, and when one
p. 87 is fully absorbed in the groove again after the break, the feel of the feel, i.e. of intensity, is gone. Put
di erently, intensity does not step forward before some kind of distanciation takes place. However,
interestingly, during this moment of experienced intensity, intensity also seems to be projected onto the
rest of the experience. Importantly, however, intensity understood in this sense does not come forward in
the form of a certain quality. The intensi cation of the right feeling does not mean that the right feeling is
being explicated or transformed into something outside itself: the right feeling is not a feeling of something.

Presence–Absence–Presence: The Shi to Meaning C5.S4

An important observation that emerges from the above is that, in the very moment one starts to consider it, C5.P35
one’s state of being is changed. This may also be viewed in reverse: there is no state of being before it is
understood and labelled as such. In this way, one’s distancing from the state of being-in-the-groove is
constitutive of the state itself.

In the case of the absorption in groove, too, a distanciation has to take place. The deferral of meaning— C5.P36
whether in the form of a ‘what has happened’ or a ‘how did it happen’ or just a ‘it has happened’—has to
come to an end. We are distanced from the situation: the event ‘goes beyond itself in the meaning’ (Ricœur
1973: 132). No account of the groove experience can escape this fact. What happens at this point, in the
moment after the event, which in the case of groove may have lasted for quite a while? The traces of the
event are still there, located within the body in the form of sensations, as traces of a rush of a certain
vitalizing energy. At this moment, we become aware of the event, or simply come to terms with it. The event
is becoming an event of understanding.
By this time, we are also regaining a consciousness of time, since when we stepped out of the time of the C5.P37
groove, temporality became situated. The temporal event is now objecti ed as an event. Now meaning
happens—now the temporality of meaning is transparent—while the temporality of the event may be
summed up as ‘time’. The event as such is distanciated, while the event as meaning is installed in time. In
other words, from a state of being characterized by an absence of meaning, it is now the event that is absent;
the presence in the event is replaced by a presence in meaning.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
In this very moment when the event has come to an end, the process of understanding the event begins, and C5.P38
the ‘quoting’ of the event may start. However, in the case of groove, an immediate resistance to inscription
seems to be considerable. Any attempt to account for the meaning of the event, embedded in the process of
distanciation, becomes meaningless, as though the traces of the event refuse to be led out of the body. They
struggle against the transition to meaning. Experiencing rhythm as presence in time, as one does in a
groove context if the groove succeeds, may be an event whose meaning is to never be meaning; meaning
dissolves in the very moment of distanciation, as if the meaning is that it (the meaning) remains within the
event, within the state of being-in-the-groove.

p. 88
Coda C5.S5

The distanciation from the state of being in the groove activates the problem of representation. The ‘it C5.P39
happens’ is impossible to represent, and this becomes crucial when facing an experience that takes place in
an almost completely absorbed mode of being. It is exactly those aspects of the event that belong to the very
being-in-the-situation—e.g. the right feeling and the feeling of feeling as such—that tend to disappear in
the shift to meaning and that lead us to qualify the event as an experience of intensity. Intensity in this
explicated form becomes the meaning of the event, while intensity as the intensi cation of a certain state-
of-being as such is hidden under the quality ‘intensity’.

The ‘it happens’ cannot be fully captured by an analytical description of the groove, even though the C5.P40
structures and features described in the analysis clearly underlie the process of getting into such a groove
state-of-being. Rather, when in such a state-of being, the technical-rhetorical mechanisms of the groove—
i.e. the typical features of rhythm that tend to keep the listener invested in the process—are transparent. In
other words, when in the groove mode of being, we do not re ect on structural tension between gures and
layers or ambivalence at the micro level of rhythm that have brought us there. As soon as we do, our mode of
being has in fact changed to an after-the-fact understanding of the event. Every attempt at re ecting on the
state of being-in-a-groove from within is thus doomed to fail; it changes what we were supposed to re ect
on.

When we are in the groove mode of being, when the ‘it happens’ actually happens, we are, on the other C5.P41
hand, completely uninterested in the fact that the main aspect of this condition, the absorbedness,
disappears in the moment that it is being ended. When playing or dancing or listening in a participatory
mode, time is not forgotten, but nor is it remembered: we are in time, we are, or rather we act. The state of
being-in-the-groove takes place exactly in that now that remains unknown to consciousness, the now that
consciousness is unable to think, the now that it even forgets in order to constitute itself.

Notes
1. It should be noted that musical rhythm can also be ʻfreeʼ—i.e. not structured according to a matrix of strong and weak C5.N1
beats, but still recognizable as such. One might also say that free rhythm establishes its own ʻmetreʼ. However, such an
understanding of metre di ers from the more traditional notion of metre as a matrix of strong and weak beats.
2. The Greek word rhythmos is related to the word rhein, ʻto flowʼ. C5.N2

3. This is the fundamental insight conveyed by Heideggerʼs Being and Time, as well as his later works ʻTime and Beingʼ and C5.N3
ʻThe Origin of the Work of Artʼ (1962; 1971, 1972). Heideggerʼs philosophy was deeply inspired by Husserlʼs The
Phenomenology of Internal Time-Consciousness, which is another classic text within this tradition (Husserl 1991).
Important philosophers following Heideggerʼs seminal works include Jacques Derrida (e.g. 2016) and Gadamerʼs Truth
and Method (1992 [1960]).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
p. 89 4. Ricœur uses the text as a paradigmatic example of the distanciation that is a part of social life as such: ʻFor me the text is C5.N4
much more than a particular case of interhuman communication, it is the paradigm of the distanciation in all
communicationʼ (Ricœur 1973: 130).

5. Derrida interprets the sign as deferred presence, in the sense that ʻthe circulation of signs defers the moment in which we C5.N5
can encounter the thing itselfʼ, so as to make us believe it is already present (1982: 9).
References C5.S6

Bengtsson, I., Gabrielsson, A., and Thorsén, S. (1969). Empirisk rytmforskning. Svensk tidsskri för musikforskning 51: 49–118. C5.P42
Google Scholar WorldCat

Butterfield, M. (2006). The power of anacrusis: Engendered feeling in groove-based musics. Music Theory Online 12(4) (Dec.). C5.P43
http://www.mtosmt.org/issues/mto.06.12.4/mto.06.12.4.butterfield.html

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Clarke, E. F. (1985). Structure and expression in rhythmic performance. In P. Howell, I. Cross, and R. West (eds), Musical structure C5.P44
and cognition, 209–236. Academic Press.
Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (1987). Categorical rhythm perception: An ecological perspective. In A. Gabrielsson (ed.), Action and perception in C5.P45
rhythm and music, 19–34. Royal Swedish Academy of Music.
Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (1989). The perception of expressive timing in music. Psychological Research 51(1): 2–9. C5.P46
Google Scholar WorldCat

Crossley-Holland, P. (2017). Rhythm. In: Encyclopædia Britannica Online. https://www.britannica.com/art/rhythm-music C5.P47


Google Scholar Google Preview WorldCat COPAC

Danielsen, A. (2006). Presence and pleasure: The funk grooves of James Brown and Parliament. Wesleyan University Press. C5.P48
Google Scholar Google Preview WorldCat COPAC

Danielsen, A. (2010). Introduction. In A. Danielsen (ed.), Musical rhythm in the age of digital reproduction, 1–19. Ashgate. C5.P49
Google Scholar Google Preview WorldCat COPAC

Danielsen A. (2015). Metrical ambiguity or microrhythmic flexibility? Analysing groove in ʻNasty Girlʼ by Destinyʼs Child. In C5.P50
R. Appen, A. Doehring, D. Helms, and A. F. Moore (eds), Song interpretation in 21st-century pop music, 53–72. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Danielsen, A. (2018). Pulse as dynamic attending: Analysing beat bin metre in neo-soul grooves. In C. Scotto, K. M. Smith, and C5.P51
J. Brackett (eds), The Routledge companion to popular music analysis: Expanding approaches, ch. 12. Routledge.
Google Scholar Google Preview WorldCat COPAC

Danielsen, A., Haugen, M. R., and Jensenius, A. R. (2015). Moving to the beat: Studying entrainment to micro-rhythmic changes in C5.P52
pulse by motion capture. Timing & Time Perception 3(1–2): 133–154.
Google Scholar WorldCat

Davies, M., Madison, G., Silva, P., and Gouyon, P. (2013). The e ect of microtiming deviations on the perception of groove in short C5.P53
rhythms. Music Perception 30(5): 497–510.
Google Scholar WorldCat

Deleuze, G. (1994). Di erence and repetition. Athlone Press. C5.P54


Google Scholar Google Preview WorldCat COPAC

Derrida, J. (1982). Margins of philosophy. University of Chicago Press. C5.P55


Google Scholar Google Preview WorldCat COPAC

Derrida, J. (2016). Heidegger: The question of being and history (trans. G. Bennington). University of Chicago Press. C5.P56
Google Scholar Google Preview WorldCat COPAC
Desain, P., and Honing, H. (1989). The quantization of musical time: A connectionist approach. Computer Music Journal 13(3): C5.P57
56–66.
Google Scholar WorldCat

Gadamer, H. (1992) [1960]. Truth and method (2nd edn, trans. J. Weinsheimer and D. G. Marshall). Crossroad. C5.P58
Google Scholar Google Preview WorldCat COPAC

p. 90 Hasty, C. F. (1997). Meter as rhythm. Oxford University Press. C5.P59

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Haugen, M. R. (2016). Music-dance: Investigating rhythm structures in Brazilian samba and Norwegian telespringar performance. C5.P60
Doctoral dissertation, University of Oslo. https://www.duo.uio.no/handle/10852/52905

Heidegger, M. (1962). Being and time. Blackwell. C5.P61


Google Scholar Google Preview WorldCat COPAC

Heidegger, M. (1971). The origin of the work of art. In Poetry, language, thought, 15–89. Harper & Row. C5.P62
Google Scholar Google Preview WorldCat COPAC

Heidegger, M. (1972). Time and being. In On time and being, 1–24. Harper & Row. C5.P63
Google Scholar Google Preview WorldCat COPAC

Husserl, E. (1991). On the phenomenology of consciousness of internal time (1893–1917) (trans. J. B. Brough). Kluwer Academic. C5.P64
Google Scholar Google Preview WorldCat COPAC

Iyer, V. (2002). Embodied mind, situated cognition, and expressive microtiming in African-American music. Music Perception C5.P65
19(3): 387–414.
Google Scholar WorldCat

Jones, M. R. (1976). Time, our lost dimension: toward a new theory of perception, attention, and memory. Psychological Review C5.P66
83(5): 323–355.
Google Scholar WorldCat

Keil, C. (1994a). Motion and feeling through music. In C. Feil and S. Feld (eds), Music grooves, 53–76. University of Chicago Press. C5.P67
Google Scholar Google Preview WorldCat COPAC

Keil, C. (1994b). Participatory discrepancies and the power of music. In C. Feil and S. Feld (eds), Music grooves, 96–108. University C5.P68
of Chicago Press.
Google Scholar Google Preview WorldCat COPAC

Keil, C. (1995). The theory of participatory discrepancies: A progress report. Ethnomusicology 39(1): 1–20. C5.P69
Google Scholar WorldCat

Kilchenmann, L., and Senn, O. (2015). Microtiming in swing and funk a ects the body movement behavior of music expert C5.P70
listeners. Frontiers in Psychology 6: art. 1232. https://doi.org/10.3389/fpsyg.2015.01232
Google Scholar WorldCat

Kvi e, T. (2004). Description of grooves and syntax/process dialectics. Studia Musicologica Norvegica 30: 54–77. C5.P71
Google Scholar WorldCat

Large, E. W., and Jones, M. R. (1999). The dynamics of attending: How people track time-varying events. Psychological Review C5.P72
106(1): 119–159.
Google Scholar WorldCat

London, J. (2012). Hearing in time: Psychological aspects of musical meter (2nd edn). Oxford University Press. C5.P73
Google Scholar Google Preview WorldCat COPAC

Matsushita, S., and Nomura, S. (2016). The asymmetrical influence of timing asynchrony of bass guitar and drum sounds on C5.P74
groove. Music Perception 34(2): 123–131.
Google Scholar WorldCat

Monson, I. T. (1996). Saying something: Jazz improvisation and interaction. University of Chicago Press. C5.P75
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469161 by National Science & Technology Library user on 26 May 2023
Nketia, J. H. K. (1974). The music of Africa. Norton. C5.P76
Google Scholar Google Preview WorldCat COPAC

Prögler, J. A. (1995). Searching for swing: Participatory discrepancies in the jazz rhythm section. Ethnomusicology 39(1): 21–54. C5.P77
Google Scholar WorldCat

Ricœur, P. (1973). The hermeneutical function of distanciation. Philosophy Today 17: 129–143. C5.P78
Google Scholar WorldCat

Rhythm. (2018). In Oxford English Dictionary. https://en.oxforddictionaries.com/definition/rhythm C5.P79


Google Scholar Google Preview WorldCat COPAC

Senn O., Kilchenmann, L., von Georgi, R., and Bullerjahn, C. (2016). The e ect of expert performance microtiming on listenersʼ C5.P80
experience of groove in swing or funk music. Frontiers in Psychology 7: 1487. https://doi.org/10.3389/fpsyg.2016.01487
Google Scholar WorldCat

Wilson, O. (1974). The significance of the relationship between Afro-American music and West African music. Black Perspective in C5.P81
Music 2(1): 3–22.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
CHAPTER

6 Politicking Musical Time  C6

Chris Stover

https://doi.org/10.1093/oxfordhb/9780190947279.013.5 Pages 91–C6.P69


Published: 08 December 2021

Abstract
Beginning with four irreducibly interrelated themes—that musical time is active, performative,
relational, and political—this chapter develops a theory of musical time as an ongoing nexus of events
that unfold between performing, listening, and sonorous bodies. It examines the temporal
implications of how these di erent body categories relate a ectively, how they co-constitute one
another, and how musicking contexts are enacted through their intra-action. The theoretical
framework draws upon Jacques Rancière’s conception of political enunciations or enactments and
Gilles Deleuze’s three syntheses of time, reading these two conceptual apparatuses productively
through one another.

Keywords: performativity, a ect, interaction, Barad, Deleuze, Rancière, Clara Schumann


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

WHAT does it mean to be ‘making time in music’? This chapter will argue for four irreducibly interrelated C6.P1
themes that together begin to clarify how music’s temporality functions. First, musical time is active. It is a
practice of actors acting, of situated bodies with perspective and agency, acting out within, and in doing so
constituting, their musical contexts. Second, it is performative. Through such actings-out, musicking bodies
perform their own subjectivities, which are in processes of being formed through those very performative
acts. The formation of musicking subjectivities is intimately implicated in the making of musical time.
Third, it is relational; what Gilles Deleuze (1978; 1988) calls ‘a ective’ and Karen Barad (2007) describes as
an intra-action that precedes and conditions the constitution of a subject. We will see that the actions of
performing, listening, and sonorously material musical bodies impinge a ectively on one another. Bodies,
therefore, are not only constituted performatively but also relationally. Implicit in this account is a
suggestion, following Barad, that non-human actors have agency. And fourth, as this chapter’s title
suggests, musical time is political, by which I will soon mean Jacques Rancière’s speci c technical de nition
of politics, as an irruption within a given social order through which democratic re-articulations are made
1
possible and political subjectivities are constituted.
Musical time is drawn through multiple processes. I mean ‘drawn’ in two ways: in the creative, artistic C6.P2
sense, as in an image of musical time being drawn as it goes along ‘on the thread of a tune’ (Deleuze and
Guattari 1987: 311), and in the sense of being drawn or pulled through a context or conceptual space, picking
up bits of genealogical residue as it goes along. It is in this double sense that musical time is made. All of this
points to a polemical proposition: that musical time is a special kind of time that occurs when, in the
context of a given social order (or what I will soon refer to as a ‘police logic’, following Rancière 1999),
musicking roles are declared and redistributed, and music’s essential excess—over and above its

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
(ac)countability—is made manifest. In other words, there are circumstances where music is taking place
that we might best regard as something other than musical time—a term that takes on a precise technical
2
meaning when we turn to a politics that engenders a speci cally musical time.

p. 92 An active, performative, relational, political account of musical time begins with bodies. To account for C6.P3
musical time in terms of bodies means, rst of all, to understand how bodies are performatively constituted:
a body’s subjectivity is performed through actions in time. As Judith Butler (1990) demonstrates, bodies
construct their identities through performative actions or gestures. Subjectivity therefore is an emergent
process: there are no ready-made subjects that go and perform those gestures; rather, subjectivity is
produced in the in-between, composed by—or inscribed through—performative acts, negotiating within
their social and historical contexts the discursive structures within which they function, and the conditions
of their materiality. For Butler, this last point refers to their biological materiality. In the case of musicking
subjectivities, we need to consider the performative actions of prosthetic-, post-, and non-human actors;
for example, the prosthetic-human materialities of performer–instrument assemblages or the sonic
materialities of musical utterances as and when they are set free from the (intra-)agential actions that
produced them and imbued them with a ective autonomy, to be taken up in relational nexuses in any
number of creative ways. I will develop these crucial points; for now, it is most important to understand
simply that ‘the forces at work in the materialization of bodies are not only social, and the bodies produced
are not all human’ (Barad 2007: 33–34).

Second, bodies exist in a ective relations with other bodies. There are many ways that we can and should C6.P4
attend to the ecological conditions within which bodies are constituted. I focus on Deleuze’s reading of
Baruch Spinoza’s a ect, which unfolds as the double movement to a ect and to be a ected. Bodies a ect and
are a ected by each other, the identity of each unfolding through impingements of multivalent a ective
forces in time. Spinoza ascribes distinct words to these trajectories: a ectus and a ectio respectively.
Deleuze evokes them in this way as well (Deleuze 1988: 49–50). When English translations collapse them
into the single term ‘a ect’, there is a loss of speci city of meaning, but we gain a sense of the irreducibility
of the two terms—they are intimately implicated in one another in a way that resonates with Barad’s intra-
3
action. I have found this irreducibility rhetorically valuable. So, building upon the performativity of bodies,
identities are constituted through a ective, temporally unfolding interactions with other bodies. Bodies
perform-for: for their own constitution and for that of other bodies. I am performing my identity; that is, I
am drawing my identity through my performed actions, passively expressing them through my next
actions. Again, I mean ‘draw’ here in the two senses already described. Those performative actions are being
taken up by the bodies with which I nd myself, and at the same time I am taking up their performed
actions. To say that bodily performativity is a ectively determined is to emphasize three key points; rst,
that performativity is social and mutually constitutive. Second, that the performativity of bodies is taken up
variably according to the di erent ways that, as Brian Massumi describes (2011: 112–114), bodies are
‘a ectively attuned’—how the ways in which the histories, social and cultural conditionings, needs, wants,
and desires of a body attune it to be a ected in a distinctive manner. And third, that bodies are both
constituted in time and participate in the constitution of temporal processes.

p. 93 Another way to say this is that bodies are ontologically temporal. This is the axiom with which all process C6.P5
philosophy begins. Bodies exist in time, express time, make time, which is to say the actions performed by
bodies in a ective relationships exist in, express, and make time. Through turning to the performative, we
learn that there is no distinction between bodies and the actions they perform—bodies are only products of
their actions; actions are only performed by bodies. The constituted-in/expressive-of/constitutive-of triad
grounds the paradoxical structure that all accounts of ‘what time is’ seem to struggle with. This is a paradox
that needs to be reckoned with by every theorist of time: how to reconcile the notion that none of these
terms is ontologically prior, with none functioning as a foundation for the other two.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
I suggested that three types of musicking bodies are at stake in the enactment of musical time: performing C6.P6
bodies, listening bodies, and musical-objects-as-bodies. Acts of listening are themselves performative acts,
so we could also conceive of two categories. Or we could extrapolate to account for many types of bodies
implicated in the time of music: composing bodies, critical bodies, sound-engineer bodies, piano-tuner
bodies, and so on. This is Christopher Small’s (1998) gambit, one political resonance of which is to
challenge conventional hierarchies that elevate certain kinds of musical subjectivities at the expense of
others, and to turn to the active doing of musicking in all of its imaginable resonances. But a trio of bodies
(performing, listening, sonorous) is a useful heuristic because it foregrounds three di erent perspectives,
each of which illuminates certain aspects of musical time. Performing bodies precede musical-objects-as-
bodies, but once the latter are enacted they become agentially autonomous. Flows of a ective relationships
proliferate in multiple directions. Listening bodies, in turn, are preceded by sonorous bodies, but as any
performer knows, listeners impinge a ectively upon the unfolding identity of a performance in ways that
can be very subtle or extraordinarily profound (e.g. the thrilling a ective interplay between musical
performer and audience during Paul Gonzalves’s famous solo in Duke Ellington’s (1956) ‘Diminuendo and
Crescendo in Blue’ at the Newport Jazz Festival). Musical meaning, therefore, emerges from complex multi-
directional a ective relationships between many types of bodies, which includes listening bodies and
performative listening acts.

What is gained by thinking of musical objects as bodies? By giving them body status, we can think of how C6.P7
the four aspects of bodies I have described—activity, performativity, relationality, and politicality—
resonate in and through musical objects and the ways in which they participate in music-temporal context-
4
constitutions. Barad insists that matter has historicity and agency (2007: 60). Sound is a kind of matter too:
the historicity of a sonic utterance is an expression of its genealogy and its context(s), and its agency resides
in its capacity to a ect and to be a ected. The yodel that triggers the alpine avalanche is agentially
decoupled from the yodeller; a song has the capacity to a ect a listener in a way that might have nothing to
do with any putative intention of the songwriter or singer. This a ect-relation is certainly conditioned by
sociocultural or ideological contexts, but not entirely so: theories of sociocultural a ordances go so far, but
there is always room for a line of escape into a new space of a ective meaning-constellations. This is one of
p. 94 the tenets of a ect theory, that we perform our attunements in individual ways that escape overarching,
structuring narratives. Musical sound has the capacity to be an equal participant in any musicking context,
which is to say any constitution of musical time.

Focusing on the third term, ‘relationality’, musical objects perform their identities through their contexts: C6.P8
this includes not only historical and syntactic contexts but also, importantly, any given temporal context in
the event of its occurring. An object like, say, a major-minor seventh chord acquires aspects of its identity
through the ways in which it interacts with, is impinged upon, and impinges upon other objects in an
ecological space that is being drawn through the very relationalities of objects that constitute it. For
example, the major triad with root a perfect fth below that follows it partially con rms its status as a
dominant chord in a key, according to a historically and culturally contingent syntactic convention.
Benjamin Boretz (1973/2003) might call the small compound structure that results, itself now a second-
order musical-object-as-body, something like a ‘tonic-followed-dominant-followed-by-dominant-
preceded-tonic’, which was a ectively implicated in the initial structure (the major-minor seventh chord
as musical-object-as-body), but only virtually: the onset of the dominant-preceded-tonic marks a
5
becoming-actual of the virtual through a selection of virtual forces to actualize, in Deleuze’s conception.
All of this, of course, also re ects (and is necessitated by) a complex array of historical and social/cultural
impingements.

Moving to the rst and second terms, ‘activity’ and ‘performativity’: much like a human (or prosthetic C6.P9
human-instrument) actor, a musical object performs its identity in order to be taken up by other objects in
a ective relationships, and in a way that is partly engendered by how it is a ected by other objects in those

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
relations. A special type of a ective relationship inheres between a performing body and a sonorous one. A
simple reading of this relationship is that musical objects are performed by bodies, and therefore that
performing bodies are prior to sonorous bodies. This is true enough—as I have described, in most musical
situations, musical objects are the sonic results of actions performed by human bodies—composing bodies,
performing bodies, computer-programming bodies, and so on. But once ‘out there’—once a virtual musical
object has become-actual through its performative enactment—it becomes part of the intra-agential social
structure through which multiplicities of identities are drawn. Once actualized, musical objects a ectly
impinge on human bodies as well as other sonorous bodies, contributing to the ongoing identity formation
of a larger musical utterance.

For example, consider a free-improvisational context: several instrumentalists performing for an audience C6.P10
of some number of listening bodies. Several minutes into the performance, I play on the trombone an
abruptly angular descending gesture with a raspy sotto voce tone, humming a line that traces a
heterophonic melodic path something like, but not quite, a semitone away from that which my vibrating
lips and slide arm are producing. This gesture abruptly and surprisingly interrupts what had been an
extended static, pointillistic, pianissimo ensemble moment. The ‘meaning’ of my now-ongoing gesture is
contingent on what has come before (far from limited to that immediate pointillistic antecedent), what is
p. 95 coming next, how various actors (performers, listeners-who-are-also-performers) constitute that
meaning (based on their a ect attunements), and how they express it for next takings-up. In the case of the
co-performing musicians, aspects of the way they take up the a ective implications of my gesture—the way
they are a ected by it—are expressed by what they do next; what next gestures they enact, and how those
gestures, immediately freed from the expressive implications of their enactment, are taken up in turn. This
is the concept that underlies what we over-simply call ‘call and response’. Any given next gesture is,
therefore, not only performative of an emergent subjecti cation: it is performative in a way that expresses
6
some aspect of the a ect impingement of the just-played gesture. My big interruptive gesture brings the
speci cally temporal implications of all of this into relief, as we will see below when we turn to Deleuze’s
third synthesis of time.

Once that gesture has been enacted, it is not only my playing of it that enters into the ongoing transaction of C6.P11
a ective forces. It is now an autonomous body, there to be a ected. Brian Massumi describes an ‘autonomy
of a ect’ that decouples e ects from their literal causes:

The body infolds the e ect of the impingement—it conserves the impingement minus the C6.P12
impinging thing, the impingement abstracted from the actual action that caused it and [the] actual
context of that action. This is a rst-order idea produced spontaneously by the body: the a ection
is immediately, spontaneously doubled by the repeatable trace of an encounter. […]The trace
determines a tendency, the potential […] for the autonomic repetition and variation of the
impingement. (2002: 31–32)

As I have suggested above, the autonomous sonorous body is also a ected by other bodies: those next C6.P13
gestures leave their a ective mark on my gesture (which is now no longer ‘mine’; was it ever?) as new
assemblages are engendered to enact new meaning-potentialities. Its identity as a body emerges through
the ways in which multiple forces impinge on it, even as it passes back into the past in light of new events.
The fourth term, ‘politics’, refers back to the title of this chapter. I suggested above (and will elaborate C6.P14
below) that politics for Rancière is enacted when there is a declaration of a fundamental equality of voices,
and an insistence that there is a part to play by that which was assumed under a hierarchical regime to have
no part. Musical sound is, under most accounts, something of an inert player in the formation of an ongoing
7
musicking context; ironically, sounded sounds have no voice! In a politics of musical time this cuts
multiple ways: I will demonstrate that musical time proceeds by way of the declaration of a part by multiple
actors that have no agency within a given logic. This agency is doubly denied, since meaningful expression

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
for Rancière is constituted at once as the possession of an expressive apparatus and as a contextual
understanding or acceptance as an entity that has such an apparatus. Rancière makes this important point
at many junctures; for example, in ‘Ten Theses on Politics’:

If there is someone you do not wish to recognize as a political being, you begin by not seeing them C6.P15
p. 96 as the bearers of politicalness, by not understanding what they say, by not hearing that it is an
utterance coming out of their mouths.

(Rancière 2001: ¶23; see also Rancièreʼs extended engagement with the relation between between phonē and logos in
the opening pages of Rancière 1999.)

The three kinds of actors I focus on here are, again, human performers, human performers, the sonorous C6.P16
8
bodies of musical utterances, and listening bodies, but again there are potentially many more.

Time ows through this nexus of constituting factors in three ways. To return to our earlier example, the C6.P17
move from a to b—from tonic-followed-dominant to dominant-preceded-tonic—takes time. Some time
passes, during which a moves to b. In this Aristotelian sense the events of a and b are in time; time is prior to
9
their enactment. Second, this move makes time in the sense that time is enacted by the movement of and
through performed events. These two modes—taking and making time—are not contradictory. Third, the
process of a moving to b expresses time in the sense, again, that it expresses a relationality not only of the
terms that constitute the event of this move, this time, but of histories of similar moves and of local and
long-range genealogies, for example what has transpired in a given performance up to the point of this
particular a → b assemblage (and what might be expected to come next). This also suggests that to express
time is to express an asymmetrical relationship of before and after: the a → b assemblage describes a and b
in terms of one another, mostly simply, a as that which comes before b and b as that which follows a. I shall
return to all of this when I turn to Deleuze’s syntheses of time. The kind of time-reckoning that I am most
concerned with here is the middle term, making time: the active, performative, relational, political doing of
musical time-constitution, but all three modes in ect one another.

The simple event of b following a—two discrete performed musical objects, one after the other, in some C6.P18
kind of syntactically and semantically grounded relationship—thus allows us to develop a basic tripartite
understanding of time as taking, making, and expressing. The active in-time gerunds are important, since
time is also something that we take up, make, and express by being caught up in the middle of the action as
productive, expressive participants. While it may seem obvious, it is important to emphasize that this
simple event is an abstraction away from even the most straightforward reality of a musicking context. In
other words, there is never a fully discrete a or b, or a single directional arrow that distinguishes one from
the other (just as there is never a discrete performative subject transcendent to the relationships that
partially constitute it). Rather, there are manifold as and bs, all impinging a ectively on one another, and
arrows proliferate wildly. The nature of those a ective impingements starts to point to a sense of a
Rancièrean politicality of musical relationality, as I have alluded to several times now, but we are not there
yet. Most important—and this is where Deleuze’s conceptual apparatus reveals its profound utility—we
should continually remind ourselves not to reduce away the complexity of time in pursuit of an elegant
formal model that in the end falsi es that which it claims to clarify.
p. 97
Deleuzeʼs (Musical) Time C6.S1

Deleuze has given us an immensely powerful conceptual model for thinking about time as three always- C6.P19
ongoing syntheses, none reducible to another, all playing a constitutive role in the movement of time’s
10
passing. Deleuze’s syntheses are passive, which means they operate at a functionally constitutive level
that precedes consciousness or cognition: as I have described (see note 5), Deleuze’s is not a theory of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
experience. But Deleuze’s conception of what I am calling the taking, making, and expressing of time can be
used to found a rich theory of experience that ows from the a ective ‘something-doing’ of time as a
11
manifold phenomenon. As passive processes, Deleuze’s syntheses go far toward explaining how a now-
ongoing context takes shape; how (active) performative gestures seize and deploy (passive) a ective force-
relations—how they contract those relations—to engender new relational expressions.

The rst synthesis of time is the foundation for time—it is where time is found (and, secondarily, where we C6.P20
experience time); in the now-ongoingness of a living present. The living present synthesizes the past and
future by contracting them into itself; the past as a swirl of passively interacting relations that condition how
any given living present can be; the future as a virtual eld from which singularities are passively selected to
become-actual. Past and future, then, are expressed in the living present; or, better, the living present
unfolds as a symptom or sign of the double impingement of active and passive selections from pure pasts and
12
virtual futures. It is important to keep in mind that there is never a single living present: multiple living
presents continuously overlap, contain, are contained within, a ect, and are a ected by one another. This
becomes evident when we think about a now-ongoing musicking context: there is the expressive living
present of my performative gesture—say, that zig-zagging trombone line—which is di erent from that of
your listening act, which is di erent from that of the monitor engineer struggling with a technical glitch or
the bored ticket-taker waiting out the clock, which is di erent from that of my performed gesture itself, in
the same movement made autonomous from my production with a new constellation of virtual past and
future relational potentials.

In Deleuze’s second synthesis of time, the past founds time. The living present is a now-ongoing expression C6.P21
of its past. But also, the living present is always already in a process of folding back, of contracting back into
past, of becoming-past. As such, it modi es the past: the past, therefore, is constantly changing. In the
terms we have been engaging thus far, the ways in which a living present is nd themselves in an a ective
relationship with the past; this relationship is, again, doubled, past and present a ecting and a ected by
one another. So while in the rst synthesis the future was shown to be a virtual eld, the past of the second
synthesis is likewise virtual.

This point bears repeating—we tend to think about the past as a thing that we have access to: it has C6.P22
happened; it is there for us. It has to be accessed through whatever interpretational apparatus we have at
p. 98 our disposal, but the sense is that we can access it, even if incompletely or incorrectly (we can
misremember, for example). This is an axiom that underlies hermeneutical projects: if we just work hard
enough, we can reveal what is there in the past to be found. In Deleuze’s reading, though, the past is a
13
virtual plane that is always already in the process of being created. Just as we enact a process of becoming-
actual of the virtual as we move into the future, the ways we can express the past in the living present re ect
the swirling of virtual forces that make up the past; there is, therefore, also a becoming-virtual of the actual
in the act of folding back into past. In other words, the past itself is being creatively drawn through those
very actions of moving into the future. This is absolutely crucial in Deleuze’s conception: the past is
creative, agentially (intra-)active and changing, and never just-there.

The time of the third synthesis is the time of the event. An event cuts time into two asymmetrical series— C6.P23
past and future. Deleuze shows how we move into the future through the enactment of events. We can do
this: I can actively enact an event that expresses a particular way of moving from past to future. Perhaps I
intended to do this with my interruptive trombone gesture. But most of the time events are just happening,
just doing, passive. And, as with the living present of the rst synthesis, there are many concurrently
ongoing events, impinging a ectively on one another. For Deleuze, an active event is a kind of conscious
selection that makes evident the eventfulness of every moment. Action, once again, is founded on passivity;
experience on the always just-stirring of the something-doing. Everything is an event. One thing Deleuze’s
third synthesis does for us, in thinking about musical interaction, is to force a turn from the agency of
conscious decision-making to the a ective genealogy of any such decision, without resorting to any sort of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
causal or deterministic narrative. In fact, the way the three syntheses come together makes this point even
more compellingly: the rst synthesis passively contracts past encounters and future possibilities and
expresses them through the enactment of a now-ongoing event; the second synthesis performs a folding of
that event into the past to newly condition future events; and the third synthesis marks the expression of
the expression, the now-ongoing gesture that cuts into and animates time.

The three syntheses of time are always-ongoing and together constitute what time is and how it is made: the C6.P24
rst synthesis as time’s foundation, the second as that upon which time is founded, the third as the engine
that propels time forward. I described this as the constituted-in—expressive-of—constitutive-of triad;
these formulations map onto the second, rst, and third synthesis respectively.

Why is this model especially compelling for understanding musical time? An example will help illuminate, C6.P25
and in doing so will set the context for a speci cally political account of musical time. Consider the opening
moment of Clara Schumann’s Three Romances for Violin and Piano (No. 1) Op. 22 (1853). The rst eleven
bars, with annotations, are shown in Figure 6.1. From the perspectives of phrase structure and metre—two
obvious locations for thinking about musical time—there is a great deal going on in this opening, all of
which opens up potential for both experiential and ontological ambiguity. There is the simple triple metre of
the piano’s opening gesture (a), rising from the midst of an inner voice to more salient prominence—a
p. 99 becoming-melodic (b). There is the double
p. 100

problematization that ensues: the violin’s entrance articulates a compound duple metre in metric
dissonance with the piano stratum (c), and also suggests an anacrusis that leaps expressively to G♭5, which
in turn sounds very much like an o -tonic phrase-beginning arrival (d). But while the violin’s arrival
engenders a beginning, thereby perspectivally recasting the opening four bars as a kind of introduction, the
piano’s continuation acts like a consequent repeat that con rms the antecedent status of what just
transpired (e). The result is what seems very much like two di erent beginnings, displaced four bars from
one another. But the violin’s beginning turns out perhaps to be no beginning at all: that G♭5 descends slowly
as one of several musical features that serve to extend the piano’s consequent phrase to six bars, with a
tonic arrival on the downbeat of the seventh bar. Throughout this passage, violin continues to articulate its
compound duple metric stratum (intensi ed by notated four-note phrase groupings that enact an
additional layer of metric dissonance (f)), piano its simple triple one. Furthermore, those rising piano
gestures seem to tumble forward rhythmically, whereas the lilting violin melody suspends any such clear
forward-directed momentum, instead occupying the untimely space of a lullaby. Two di erent kinds of
momentum are engendered therefore, corresponding to the two metric strata.
Figure 6.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
Clara Schumannʼs Three Romances for Violin and Piano No. 1 Op. 22 (1853). First eleven bars with annotations. C6.F1

All of this operates as a nexus of contexts for productions of certain kinds of musical times, lters for or C6.P26
limits on ontogenic possibilities—or, if we take seriously the a ective potential of non-human sonorous
bodies, the a ect attunement of the music-as-body, prior to the connections of performative enactments.
This is all part of the genealogy of the event of making musical time—part of the past of which the present
of any given performative gesture is the most contracted aspect. All of these contexts function as part of the
fabric that conditions how a given performance will transpire. There are many more contexts, of course:
historical contingencies, stylistic norms, acoustic considerations of violin–piano assemblages, and on and
on. As the second synthesis of time demonstrates, that past is also in ux, ever newly impinged upon by
next events, which engender new pasts in ongoing series. For example, a new analytic insight can in ect old
knowledge, causing one to think di erently about one’s own earlier experiences.

These contexts are also contracted into the living present of the rst synthesis. This means that, in the C6.P27
living present, passive selections are made that actualize aspects of the pure past, that bring them to life in a
particular way. In musical contexts, those selections are made by performers; the way that Nicholas Cook
frames this is by asking us to ‘think of performers as creating meaning within the structural a ordances of
compositions’ (Cook 2013: 68). Note that this selection can be made by a performer but also by a listener as a
form of hearing-as; say, ltered through the context of a particular listening strategy. In either case, the
passive or active selection that creates meaning is a performative action that enacts the cut of the third
14
synthesis, and it is this cut that moves time forward.

To say that performers create meaning by operating within the structural a ordances of their contexts is to C6.P28
15
begin to turn to the political. In this way we can also start to examine what it even means to create
p. 101 meaning, which I will suggest unfolds as a process that at least partially resides in the making of musical
time as a Rancièrean politics, to which I now turn.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
A Politics of Musical Time C6.S2

In the brief example that we just encountered, I have begun to lay out a few of the many empirical-syntactic C6.P29
factors that come together to partially condition the temporally unfolding event of any given performance
of Schumann’s Romance. One implication here is that these factors should somehow be attended to; indeed,
they must be attended to if a performance is deemed to be successful according to the logic of the social
order within which Schumann’s music locates, or in order for the event of any given living present to
contract its pasts (and futures) into it to make a musical expression that works within that logic. That social
order, policed at once by no one in particular and by everyone who operates within it, draws upon
performance practice tradition, normalized melodic-harmonic-rhythmic syntax, conventions of space and
place and decorum, concepts of beauty and expressiveness, and social or labour roles within a hierarchical
structure in order. All of these, then, are contexts within which musical time unfolds—or, read another way,
constraints on how musical time can unfold.

Jacques Rancière’s political thought begins within what he calls the ‘partition’ or ‘distribution of the C6.P30
sensible’. Brie y, he wants to nd ways to expose how the sensible (e.g. what is possible to hear, say, see,
do, etc.) is distributed, by whom, and for what reasons. He refers to the structures and forces that enact
those distributions, which require and express a logic of inequality, a ‘police logic’, which ‘arranges reality,
in the sense that it distributes people and things into locations and roles’ (Chambers 2013`; 70; see also
Rancière 1999: 29). Understanding this is a rst, necessary condition for acting in ways that might disrupt
these distributions. These are political actions. In other words, politics, as a disruption of hierarchical
distributions of what can be said and heard, or an act of dissensus, is staged within the ongoing context of the
police logic as an expression of equality that forces an awareness of the ‘wrongness’ of the police logic’s
claim of hierarchical inequality. More precisely, it does this by enacting a wrong or a miscount, a ‘count of
the uncounted’ (Rancière 2004: 13). As Rancière insists, ‘politics does not stem from a place outside the
police. […] There is no place outside of the police. But there are con icting ways of doing things with the
“places” that it allocates: of relocating, reshaping, or redoubling them’ (Rancière 2011: 6). As Samuel
Chambers glosses, ‘the “wrongness” that politics asserts in the face of a police order is also a
“wrungness”—a twisting or torsion of the police order and its logic of inequality’ (Chambers 2013: 57–58);
in other words, not only is politics grounded on the structural apparatus of the very police count that it
subverts, it acts upon and somehow transforms that structural apparatus through the event of its staging. The
metaphor of wringing or twisting will become valuable when we begin to home in on what exactly the
p. 102 musical time is that is politicked, and how that politicking is staged. Note that the temporality of these
terms—‘staging’, ‘enacting’, ‘wringing’, and ‘twisting’—all refer to the eventfulness of any given political
action.

In music, a distribution of the sensible might be a claim of syntactic structure, a discipline-endorsed music- C6.P31
analytic framework, a learned historicized constraint on performance practice, or a critical orientation that
privileges form over matter. It can be a conception of metre as a particular kind of hierarchical framework, a
claim that twelve (equal-tempered) tones are ‘only related to each other’, an assertion of chord–scale
isography, or a socially inscribed set of constraints about what constitutes a ‘beautiful’ or ‘correct’ clarinet
tone. It can also be a hierarchical division of roles and identities. Rancière describes how identities are
assigned in a police count, in doing so revealing the speci c problematic—that of the assignation of a part
to one who has no say in the assignation—that politics announces and that it must engage,

In politics ‘woman’ is the subject of experience—the denatured, defeminized subject—that C6.P32


measures the gap between an acknowledged part […] and a having no part. ‘Worker’ […] is similarly
the subject that measures the gap between the part of work as social function and the having no

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
part of those who carry it out within the de nition of the common of the community. […] Any
subject is a disidenti cation, removal from the naturalness of a place, the opening up of a subject
space where anyone can be counted […]

(Rancière 1999: 36)

Subjects, for Rancière as for Deleuze and Guattari, Butler, and many other post-structuralist thinkers, are C6.P33
not given but made through performative processes within determining contexts, while also being in
agonistic relationships with them. The state apparatus for Deleuze and Guattari, systems of power and
discourse for Butler (and, of course, Foucault), are police logics for Rancière. The process of subjectivation
for Rancière is one of disidenti cation with one’s role within a given count; the assertion of a voice within a
structure that denies such an assertion. This is, rst of all, an insistence that a radical equality has been
covered or suppressed by the hierarchical inequality of the social order that assigns places and identities to
all. To assert a radical equality is also to reveal the contingency of the police order, to demonstrate, through
16
action, that that order is arbitrary rather than somehow ordained by a natural or divine logic. Importantly,
for Rancière, subjects are only formed through practices of disidenti cation.

A Rancièrean musicking subject would be produced similarly. In the police logic of musicking-identity C6.P34
ascriptions (focusing for a moment on composed, score-availing music), the performer has a role, which is
to interpret the visual information on the page according to a set of social norms agreed upon by a given
micro-community. The performer’s role is to animate or vivify the composer’s music, to bring it to life, to
make it music through its temporalization. According to the police logic, the performer has a part in the
hierarchical order of things; likewise the composer, who seemingly remains in the background but really
occupies the space at the top of the hierarchical order: these are her notes to animate, this is her vision to
p. 103 realize. The performer, under this count, is mute, is assigned a representational task, which is not to say
that performers simply mimic, but rather that they produce their musical utterances within a regime that
determines proper ways of doing so. For Rancière, most artistic expression occurs within a representational
regime that legislates the relationship between poiesis (a ‘way of doing’) and aisthesis (‘a way of being
a ected by it’; Rancière 2013: 7), determining proper ways of doing and of being a ected by. This amounts to
legislating how one can be performative, how one can be relational, how one can (or ought to) be a ectively
17
attuned. The representational regime, then, corresponds to Rancière’s police logic. The action that
disrupts and redistributes this legislative process is what Rancière calls ‘aesthetic’, which aligns so precisely
with the irruption of politics in the police logic that Rancière often refers to a ‘politics of aesthetics’ and an
18
‘aesthetics of politics’. Importantly, both politics and aesthetics are staged within the context of the very
social order they disrupt; these are not, then, oppositional dialectical moves, but what Deleuze would call
19
‘di erential’ ones, transformative of the spaces they a ect without negating them. ‘It is within the
mimetic [police] regime that the old stands in contrast with the new. In the aesthetic [political] regime of
art, the future of art […] incessantly restages the past’ (Rancière 2004: 24).

A political (or aesthetic) music enactment occurs when a performer, through the action of a musical C6.P35
utterance, declares their equality in the order of the work-logic. Through a declaration of equality,
performers constitute their subjectivities as musicking participants by disidentifying with the police count
that establishes and continually rea rms the hierarchy of roles in the social order of music-making
contexts. It is important to clarify that a police logic is not necessarily ‘bad’, oppressive, or a problematic
structure to be overcome. Politics, again, does not exist in contrast to or in negation of police logics.
Rancière makes this very clear, insisting that his usage is ‘ “neutral”, nonpejorative’ (1999: 29), and
carefully arguing:

there is a worse and better police—the better one, incidentally, not being the one that adheres to C6.P36
the supposedly natural order of society or the science of legislators, but the one that all the
breaking and entering perpetuated by egalitarian logic has most often jolted out of its ‘natural’

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
logic. (1999: 31)

A political action, then, reveals the contingency of the police logic, reveals it to be a miscount. The C6.P37
declaration of a wrong that enacts a political miscount takes place within, disturbs, and redistributes the
logic of the existing count, thereby transforming it. This involves, once again, a disidenti cation that
relocates, reshapes, or redoubles the terms of the count. This redistribution—or reterritorialization, to fold
Deleuze and Guattari back into the picture—is, importantly, in excess of the count. It does not simply
engender (or sediment or molarize into) a new count: in Rancière’s words, it reveals the impossibility of the
count, the count that cannot be counted. Politics always exceeds the count(able).

All of this has profound music-temporal implications. My argument here is that musical time is nothing C6.P38
p. 104 less than the time of the production of (active, performative, relational, and political) musicking
subjects. As a Rancièrean politics, this production begins with the declaration of equality, of having a part in
a police logic that o ers no such part. This in turn transforms the very logic that it interrupts. Let us return
to the Schumann example to engage a more concrete illustration.

The empirical-syntactic contexts I have described are signi cant only to the extent that they constitute C6.P39
actionable aspects of the genealogy of any given now-ongoing event—that is, to the extent that they form a
part of what any such event expresses. As Deleuze makes clear, everything is an event; but let us for now
focus on the event of a violinist and pianist performing this opening moment. As interpreters of the score-
script before them, they actively and passively fold a multiplicity of contexts into the now-ongoing
performative moment. These contexts include piece-speci c information such as that described, and also
broader implications of genre norms, their own prosthetic instrumentalities engendered by years of careful
practice, histories of encounters with this work and others like it (and unlike it), teachers’ suggestions and
admonitions, spaces and places of previous encounters, and much, much more. All of this is expressed in
Deleuze’s rst synthesis of time, as are the openings onto possible futures (e.g. how this performance is
going to unfold temporally) that are made at least partially determinable through the constraints of one’s
singular past. That past is, too, in a process of changing as each now-present event folds back into it,
reconstituting it as ever-new. The moment of a next event has a di erent genealogy that the moment of the
just-past one. The events described here are the cuts of the third synthesis of time: the now-ongoing, acted,
performative, relational events that drive time forward and through which we come to understand that, in
and through every event, nothing will ever be the same again.

What makes a speci cally music-temporal event? What makes a musical event a political irruption in the C6.P40
police count of a social order? In the limited scenario of two performers staging the opening of Schumann’s
Romances, it is the declaration of equality that is made when the performers enact a miscount which
destabilizes the regime that determines what is or is not sensible and how. It is how, through their
performative acts, they disidentify with their mute roles in the police logic. The way in which the performers
use expressive microtiming is a good case study.

The temporal pushes and pulls of expressive microtiming operate on something like a normative or C6.P41
idealized version of a musical work’s temporal unfolding. This is one reason Charles Keil (1994) can invoke
a term like ‘participatory discrepancies’, the implication being that there is a norm against which
discrepancies may be measured. This norm is imaginary in an important sense, but it also forms part of the
police logic of ‘measured’ music: whatever de nition of musical metre we choose to cleave to, there is at
least a shared assumption that metre measures something, and in an overwhelming majority of contexts,
the things that are measured (whether beats, pulses, grids, etc.) are what Justin London calls ‘nominally
isochronous’ (2012: 72), meaning that we can relatively unproblematically describe them as being periodic.
Played events redistribute the logic of a nominally isochronous periodic grid. These redistributions
communicate a ectively: next played events communicate with earlier ones: those earlier events, decoupled
p. 105 from the agential forces that created them, become autonomous bodies, each a locus of and for a ective

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
forces, each participating in the constitution of further contexts. This is what I mean by performative
engagement with music’s sonorous materiality: each new event unfolds by re-relating a ectively with
earlier and ongoing (and virtual future) contexts to enact a process of subjectivation through
disidenti cation. That much of this occurs at the pre-conscious a ective level does not deny its agency, as
I’ve explained.

Expressive micro-timing unfolds in an a ective relationship with the pure past of a given event. This C6.P42
includes the syntactic sca olding of the music, as interpreted through any number of analytic lters, and
including (but far from limited to) the brief roster described in relation to Schumann’s Romance. For
example, a violinist may linger on the third B♭4 of bar 4, enacting an expressive agogic action slightly
delaying the leap up to the phrase-initiating G♭5. This is what happens in Sini Simonen and Tanya
Zapolska’s recent performance; in fact, violinist Simonen slows down through the second half of the bar
20
such that her phrase stretches noticeably out of alignment with Zapolska’s piano stratum. This has the
e ect of perceptually amplifying the distance between metric strata (compound duple for Simonen, simple
triple for Zapolska)—a process that becomes more exaggerated four bars later when Simonen’s descending
melody lags consistently behind that of her co-creator. Zapolska in turn lingers on her nal bar-4 F/A♭ dyad
to begin bar 5 in line with Simonen’s downbeat arrival; her ensuing phrase, consequent to her opening
antecedent (and in its enactment establishing her opening melodic gesture as an antecedent), stretching
and contracting much like the earlier phrase did: lingering on her initial bass onset, pushing through the
middle of the bar as piano-accompaniment texture becomes-melody for the second time. All of this is in the
process of constituting a context—the context that partially conditions each next event and that, as a
political or aesthetic irruption, contributes to the emergent subjectivity of the performer–instrument–
sound assemblage now ongoing. Why is this political? Why is it an example of politicking musical time (and
therefore why is it constitutive of a speci cally musical time)? Because, through the actions of their active
and passive selections—through the ways they perform their subjectivities in relation with nexuses of
pasts, virtual futures, and co-occurring presents—they declare their participation in the constitution of
musical meaning, which is revealed as not simply given but made, made through the temporal enactment of
musicking acts. Simonen’s and Zapolska’s performative utterances take time in that they follow the order of
the virtual events of the score-script, in communication with a long performance practice lineage and their
own singular pasts. But much more importantly, they make time in that their multi-relational utterances,
each an expression of time as a cut that assembles time, enact the time of this singular performance, as a
continuous ow of events that cut into time and assemble new pasts and futures into new relations, newly
contracted in new living presents. As a listener, I make time too, in a ective dialogue with the performance:
with the event of each new engagement I disidentify with the count of my now-past listening self
(conditioned by my expectations about what can be) while newly constituting my musicking subjectivity by
folding new events into the what-is-possible of my understanding of musical expression, which is to say of
musical time.

p. 106 Contrary to Rancière’s insistence about the frequency of political occurrences, musical time is not rare: it C6.P43
can erupt anywhere and does so frequently. Pablo Casals’s expressive ‘leading tones that lead’
deterritorialize the tempered substrate of the music he interprets. Miles Davis continually redistributes the
logic that determines what does or does not count as a good trumpet tone. Brazilian concert-going
audiences declare a redistribution of musicking roles every time they join in to sing along with a favourite
artist. Performers across many musical contexts push and pull their performed gestures to enact complex
temporal relations with other co-occurring musical strata. The eruption of a musical politics—of musical
time—is not rare, but it always occurs in relation to a police logic that suggests that there are ‘proper’ ways
of taking, making and expressing time. To rethink musical time as a (quasi-)Rancièrean politics is to
imagine a di erential logic whereby acts of dissensus continually disrupt the order of a given logic, but one
that is continually being remade in order to enact new distributions. Slightly contrary to what I have already
suggested, this does then operate like Deleuze and Guattari’s double movement of territorialization and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
reterritorialization—always ongoing, productive of its contexts; every deterritorializing breakaway co-
21
creating, in its very enactment, with a reterritorializing redrawing of new boundaries. It also resonates
with Deleuze’s manifold staging of his three interwoven syntheses of time, each of which singularly
expresses the performative, relational actions that constitute the movement of time. A common, crucial
thread that connects all of these perspectives is that the subjectivity of any given participant is composed
through processes of disidenti cation and re-relation. These processes are largely passive, and when they
are (or seem) active, as in the announcement of Rancière’s political eruption, it is an activity grounded on
passivity, as an expression of passive forces in the second synthesis of time, mobilized into disidenti catory
practices in every given temporal event of the third synthesis. All of this is to say that time is not a thing, or
a place, it is an ongoing nexus of events. Musical time, as a practice that brings time’s relational
constitution into relief, is a process involving performative actions from, for, and between human,
prosthetic-human, and sonorous musicking bodies. It is for this reason that studying musical time, as an
enacted form of politics, has extraordinary implications for research far beyond music studies.

Notes
1. This means that I will draw a connection between a musicking subjectivity and a political one, an a inity that will emerge C6.N1
in the last part of this chapter. This is not an especially radical move in terms of Rancièreʼs philosophy: since politics for
Rancière involves redistributions of sensoria, its enactment is fundamentally aesthetic (see Rancière 2004: 3). Aesthetics
(including musical aesthetics) and politics articulate similar types of concept spaces. I will, eventually, also read against
Rancièreʼs claim that politics is something that ʻactually happens very little or rarelyʼ (Rancière 1999: 17) when I locate
politics alongside the cut of Deleuzeʼs (1994) third synthesis of time and the double movement of Deleuze and Guattariʼs
p. 107 (1987) territorialization and deterritorialization. In both of the latter cases, what is originally shown as a special or
radical moment is soon revealed to be in a process of taking place at all times, multiply and necessarily.

2. There are a inities with this conception of musical time and Rancièreʼs aesthetic regime of art (Rancière 2009: 8); most C6.N2
important, how enactments of musical time call into question the contingencies of the partitions of the representative or
mimetic regime. Iʼll return to this in the last part of this chapter.

3. I develop this point in Stover 2018a—see note 21. C6.N3

4. Baradʼs non-human agency is ultimately revealed to be a participant in a process of intra-action, which foregrounds the C6.N4
relational nature of all agency, including that which we can describe in conventional terms. Barad writes:

Crucially, agency is a matter of intra-acting: it is an enactment, not something that someone or something has.
[…] Agency is ʻdoingʼ or ʻbeingʼ in its intra-activity[… It] is about changing possibilities of change entailed in
reconfiguring material-discursive apparatuses of bodily production, including the boundary articulations and
exclusions that are marked by those practices in the enactment of a causal structure. (2007: 178; emphasis
original)

Baradʼs conception resonates with both Deleuze and Rancière in at least three important ways. First, in calling into
question the location and means of productions of agency—in both cases in the active in-between, through encounters
and events. Second, by insisting that subjects (and, for Barad, objects) are constituted through those intra-active
encounters. And third—and this is where Baradʼs thought aligns with Rancièreʼs politics—in showing how these
encounters continually reconfigure the boundaries and partitions of discursive practices, a point that I will read in musical
terms, in order to carefully reconfigure a few aspects of Rancièreʼs political theory, below.
5. For Deleuze, the virtual is also real, which is to say that the implications of the constitution of the compound structure C6.N5
tonic-followed-dominant-followed-by-dominant-preceded-tonic were present prior to the onset of the second chord, in
the sense of their potential to be actualized. This potential is related to experience and expectation, but only to an extent:
Deleuzeʼs theory is not phenomenological (and therefore the future-as-virtual is not a form of protension), although
experience and the experience-able are certainly among its resonances. To say that the movement of dominant to tonic is
an actualization of virtual forces is to define the actual dominant, now in progress, in terms of its relationality to all of its
virtual implications: virtual, therefore, is not an opening onto an imminent future but refers to the e ects of that opening
on the now-ongoing event of the dominant. Furthermore, virtual is not merely future-directed: the past is virtual as well in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
the sense that it is unfixed, ever changing in light of new presents folding back into it, a ecting and a ected by it.

6. Note the language here: not just the a ect attunement of the player of the next gesture, but of the intra-agential gesture C6.N6
itself. This is of course a controversial claim, and I am glossing it slightly to make a rhetorical point, but it is a foundational
principle of new-materialist thought.

7. While this is not the place for an extended engagement with the a ect of acousmatic contexts, I would argue that, since C6.N7
such contexts radically a irm the body status of a given sound as decoupled from the agencies that created it, they should
prove to be fruitful spaces for continued inquiry in the directions I propose here. In other words, acousmatic bodies,
already decoupled from their performed sources, might be thought of as agential in ways that are even more obvious than
conventionally performed ones.

p. 108 8. This formulation draws on Smallʼs (1998) concept of musicking, with a new-materialist turn. Small insists that all number C6.N8
of di erent kinds of participating bodies can be said ʻto musicʼ in the constitution of any given performative context. This
might include stagehands, piano tuners, ticket-takers, journalists, record company personnel, club owners, teachers,
theorists, and many, many more body types (positing any given body as, in large part, a function of its performing-as). The
connection between Small and Butlerʼs theory of performativity is fruitful here.

9. Of course, Aristotle tries to reduce time to the measure of motion, which is not what I mean here: the ʻmovementʼ of a to b C6.N9
should be thought of as purely metaphorical.

10. The three syntheses of time are presented in ch. 2 of Di erence and Repetition (Deleuze 1994: 72–91). I lay out the C6.N10
foundation for a Deleuzian conception of musical time in Stover (2017).

11. Iʼm eliding a lot of complex information here, including Deleuzeʼs pre-conscious ʻdark precursorʼ (Deleuze 1994: 119), C6.N11
adapted from Henri Bergsonʼs (1990: 39) ʻzone of indeterminationʼ, which precedes cognition and responsive action. Brian
Massumi draws a connection with William James when he invokes a ʻsomething doing that is always already just stirringʼ
(Massumi 2011: 27; see also James 1996: 161). See Stover (2018b: 150–151) for more on the elaboration both of the
musical implications of these concepts and of the role music plays in understanding this aspect of temporal process.

12. See Deleuze (1983: 75–79) for a development of Nietzscheʼs conception of the philosopher as semiotician or C6.N12
symptomatologist. From this perspective, the musicologist acts as a kind of symptomatologist too, whose task is, for
example, to read to read the expressions of past and future impingements in a given musical living present.

13. The Deleuzo-Guattarian plane is never just-there, it is always in a process of being created through co-occurring acts of C6.N13
territorialization and deterritorialization. See, for example, Deleuze and Guattari (1987: 70): ʻthe plane of consistency is
occupied, drawn by the abstract Machine; the abstract Machine exists simultaneously developed on the destratified plane
it draws, and enveloped in each stratum whose unity of composition it defines. […] That which […] dances upon the plane
of consistency thus carries with it the aura of its stratum, an undulation, a memory or tension. The plane of consistency
retains just enough of the strata to extract from them variables that operate in the plane of consistency as its own
function.ʼ To keep relatively consistent with the language I have been using thus far, we can read ʻabstract Machineʼ as ʻco-
occurring processes of territorialization and deterritorializationʼ, and ʻstratumʼ likewise as the image of the first term in
that process, the appearance of a structuring that results from a territorializing act.

14. I have been emphasizing passive synthesis throughout this passage, in order, again, to steer away from a conception of C6.N14
musical time-constitution that centers around agency. Another way to say this, which brings back my earlier point about
the potential agency of non-human actors (especially sonorously material musical bodies), is to suggest that agency need
not be active or conscious. To suggest that the way I perform a musical action is a product of the passive synthesis of
many di erent kinds of prior actions, some conscious, many not, is not to eschew agency or to suggest some kind of
psychological determinism, but to construct a concept of agency that begins with the passive flows of a ective forces that
express earlier actions by (passively) seizing and creatively redeploying them. So when that sceptical young virtuoso who
appears in many a first-year theory course worries, for example, about the about the putative analytic stranglehold that
p. 109 music theory has on the free flow of musical expression (ʻparalysis by analysisʼ; ʻforget about all that and just playʼ), it
is helpful to keep in mind that every analytic act folds back into the past (in the second synthesis of time) to become part
of the genealogy of any later performative act, which passively draws the earlier gesture into itself (in the first synthesis).

15. I slightly modify Cookʼs formulation by eschewing the word ʻcompositionʼ as he intends it, since compositions are C6.N15
obviously not the only contexts that engender di erent kinds of a ordances within which to create meaning. I should also

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
push against the word ʻstructuralʼ: there are certainly many kinds of a ordances available to performers besides structural
ones.

16. See Rancière (2010: 50–51). One such contingency-masked-as-natural-law is the long history of appeals to the harmonic C6.N16
series to validate certain kinds of tonal systems over others.

17. To shi our focus to the performative, relational listenerʼs perspective, this also means legislating modes of hearing, such C6.N17
as the kinds of ʻhearing-asʼ legislated by music appreciation curricula.

18. Rancière (2004). Throughout his writings on aesthetics, Rancière a irms artʼs status as an event that transforms sensible C6.N18
experience in order to continually reshape experience as art. See e.g. Rancière (2014: ix–x).

19. One key di erence between Deleuze and Guattariʼs and Rancièreʼs conceptions is that for the former, actions like C6.N19
deterritorialization, destratification, molecularization, and so on, which might be said to resonate with the irruption of
Rancièrean politics, are always-ongoing—there is no act of deterritorialization without a corresponding reterritorialization
(see, for example, Deleuze and Guattari 1987: 218–220)—whereas for Rancière political (or aesthetic) acts are quite rare. As
I suggested above, one of my aims in bringing Deleuzian and Rancièrean concepts together is to propose a way of thinking
through the latter such that, in certain circumstances, politics, as a declaration of equality that reconfigures the sensible,
is always-ongoing as well, in the sense that it motivates a continual escape from frame through which we continually re-
engage the distribution of the police count, now as a lively process of becoming rather than a manifold series of
foreclosing gestures.

20. https://www.youtube.com/watch?v=yJFcJOFwtE4 (accessed 3 Sept. 2021). C6.N20

21. This formulation draws on Deleuze and Guattariʼs evocation of how the ritornello operates; see Deleuze and Guattari C6.N21
(1987: 311–312).
References C6.S3

Barad, K. (2007). Meeting the universe halfway: Quantum physics and the entanglement of matter and meaning. Duke University C6.P44
Press.
Google Scholar Google Preview WorldCat COPAC

Boretz, B. (1973/2003). What lingers on (, when the song is ended). In J. K. Randall and B. Boretz (eds), Being about music: C6.P45

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
Textworks 1960–2003, vol. 1, 421–428. Open Space.
Google Scholar Google Preview WorldCat COPAC

Butler, J. (1990). Gender trouble: Feminism and the subversion of identity. Routledge. C6.P46
Google Scholar Google Preview WorldCat COPAC

Chambers, S. (2013). The lessons of Rancière. Oxford University Press. C6.P47


Google Scholar Google Preview WorldCat COPAC

Cook, N. (2013). Beyond the score: Music as performance. Oxford University Press. C6.P48
Google Scholar Google Preview WorldCat COPAC

Deleuze, G. (1978). Lecture transcripts on Spinozaʼs concept of a ect. https://www.webdeleuze.com/textes/14 C6.P49


Google Scholar Google Preview WorldCat COPAC

Deleuze, G. (1983). Nietzsche and philosophy (trans. H. Tomlinson). Continuum. C6.P50


Google Scholar Google Preview WorldCat COPAC

Deleuze, G. (1988). Spinoza: Practical philosophy (trans. R. Hurley). City Lights. C6.P51
Google Scholar Google Preview WorldCat COPAC

p. 110 Deleuze, G. (1994). Di erence and repetition (trans. P. Patton). Columbia University Press. C6.P52
Google Scholar Google Preview WorldCat COPAC

Deleuze, G., and Guattari, F. (1987). A thousand plateaus: Capitalism and schizophrenia (trans. B. Massumi). University of C6.P53
Minnesota Press.
Google Scholar Google Preview WorldCat COPAC

Ellington, D. (1956). Diminuendo and Crescendo in Blue. On Ellington at Newport: Complete. Columbia Records.
Google Scholar Google Preview WorldCat COPAC

James, W. (1996) [1912]. Essays in radical empiricism. University of Nebraska Press. C6.P54
Google Scholar Google Preview WorldCat COPAC

Keil, C. (1994). Participatory discrepancies and the power of music. In C. Keil and S. Feld (eds), Music grooves: Essays and C6.P55
dialogues. University of Chicago Press.
Google Scholar Google Preview WorldCat COPAC

London, J. (2012). Hearing in time: Psychological aspects of musical meter (2nd edn). Oxford University Press. C6.P56
Google Scholar Google Preview WorldCat COPAC

Massumi, B. (2002). Parables for the virtual: Movement, a ect, sensation. Duke University Press. C6.P57
Google Scholar Google Preview WorldCat COPAC

Massumi, B. (2011). Semblance and event: Activist philosophy and the occurent arts. MIT Press. C6.P58
Google Scholar Google Preview WorldCat COPAC
Rancière, J. (1999). Disagreement: Politics and philosophy (trans. J. Rose). University of Minnesota Press. C6.P59
Google Scholar Google Preview WorldCat COPAC

Rancière, J. (2001). Ten theses on politics (trans. R. Bowlby and D. Panagia). Theory & Event 5(3). C6.P60
Google Scholar WorldCat

Rancière, J. (2004). The politics of aesthetics: The distribution of the sensible (trans. G. Rockhill). Continuum. C6.P61
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469239 by National Science & Technology Library user on 26 May 2023
Rancière, J. (2009). Aesthetics and its discontents (trans. S. Corcoran). Polity Press. C6.P62
Google Scholar Google Preview WorldCat COPAC

Rancière, J. (2010). Dissensus: On politics and aesthetics (trans. S. Corcoran). Continuum. C6.P63
Google Scholar Google Preview WorldCat COPAC

Rancière, J. (2011). ʻThe thinking of dissensus: politics and aestheticsʼ. In Paul Bowman and Richard Stamp (eds), Reading C6.P64
Rancière, 1–17. Continuum.
Google Scholar Google Preview WorldCat COPAC

Rancière, J. (2013). Aisthesis: Scenes from the aesthetic regime of art (trans. Z. Paul). Verso. C6.P65
Google Scholar Google Preview WorldCat COPAC

Small, C. (1998). Musicking: The meanings of performing and listening. Wesleyan University Press. C6.P66
Google Scholar Google Preview WorldCat COPAC

Stover, C. (2017). Time, territorialization and improvisational spaces. Music Theory Online 23(1). C6.P67
https://mtosmt.org/issues/mto.17.23.1/mto.17.23.1.stover.html
Google Scholar WorldCat

Stover, C. (2018a). A ect and improvising bodies. Perspectives of New Music 55(2): 5–66. C6.P68
Google Scholar WorldCat

Stover, C. (2018b). A ect, play, and becoming-musicking. In M. Bohlmann and A. C. Hickey-Moody (eds), Deleuze and children, C6.P69
145–161. Edinburgh University Press.
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
CHAPTER

7 To Be in
C7 Time: Repetition, Temporality, and the Musical

Work 
Nathan Mercieca

https://doi.org/10.1093/oxfordhb/9780190947279.013.6 Pages 111–C7.P70


Published: 08 December 2021

Abstract
This chapter begins by examining recent scholarship in ‘music as performance’, especially that of
Nicholas Cook, and its implications for the work concept. By exploring various formulations of the
work concept from a temporal perspective, it becomes clear that contradictions occur whenever the
work concept is tied too closely to the notion of a musical work’s identity. Instead, a Deleuzian
understanding of the musical work is advanced, based on Deleuze’s idea of repetition: this is seen as
allying closely with a deconstructive approach to musical material, which provides an additional
opportunity to consider musical temporality, in the arena of history and the musical past. Finally, to
recapture the spirit of Cook’s original theories, and drawing on Hannah Arendt, a parallel between
musical and human ontology is drawn, based on their identical interaction with time, which
reconstitutes but fundamentally changes the idea of (the) musical ‘work’.

Keywords: temporality, work-concept, ontology, deconstruction, Nicholas Cook, Deleuze, Arendt, jazz.
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Performance and the Musical Work: Spatiality or Competing C7.S1

Temporalities?

IMPLICIT in Nicholas Cook’s theories of musical performance—an important body of research spanning 15 C7.P1
years—is an inner con ict concerning the ontology of music: speci cally, the nature of the musical ‘work’,
1
and its relationship to music as a sounding phenomenon. He forcefully rebu s a common mid-century

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
formalist understanding: ‘the approach to performance epitomized by Wallace Berry understood the score,
or perhaps more precisely the structure embodied in the score, to be the locus of musical meaning. […]
Analytical understanding was assumed to be the foundation of articulate performance, and the result was to
place the theorist in a position of authority’ (Cook 2012a, unpaginated: para. 8). His distaste for this
approach is uncontroversial enough, and given the revolution in musicological thought at the turn of the
last century, I will not dwell on the sound reasons for dismissing such a narrow view of musical meaning. It
is what John Rink terms a ‘prescriptive’ mode of music-making, in which the musical work gures as a
purely conceptual phenomenon which can be (partially) projected in performance, aided perhaps by
2
theoretical and analytical tools.

Rejecting this Platonic, ‘spatial’ view of the musical work—whereby the work undergirds a performance C7.P2
(which presents only a fragment of it) or is located in one particular element of it (e.g. a score)—Cook has
posited an alternative ontology with decidedly temporal concerns. He suggests that ‘a given performance of
p. 112 Beethoven’s Ninth Symphony, for example, will acquire its meaning from its relation to the horizon of
expectations established by other performances. […] There is no ontological distinction among the di erent
modes of a work’s existence, its di erent instantiations, because there is no original’ (Cook 2012b: 187). In a
decidedly Deleuzian vein, Cook concludes that ‘instead of a single work located “vertically” in relation to its
performances, we have an unlimited number of instantiations, all on the same “horizontal” plane’ (p. 187).
Although invoking a spatial metaphor, Cook’s underlying conceptualization is clearly temporal: the
horizontal plane to which he alludes is the onward march of time itself.

On the other hand, Cook has also advocated an entirely di erent model for the musical work. Re ecting his C7.P3
interest in the burgeoning eld of performance studies, his thoughts resemble more closely what John Rink
has termed a ‘descriptive’ outlook: one which accounts not only for the score-based elements of musical
performances, but for those elements previously considered peripheral, such as spontaneous rubato in
performance, gesture, and bodily movement, and the clothing of the performers. As I shall explain, this
model fundamentally a ects the putative ontology of the musical work, and Cook is not alone in advocating
it. Linda Dusman has written that ‘performance [is] a temporary community in which composers,
performers, and audience members are all active participants, with an audience’s hearing of a work as the
nal step in its creation. In other words, music does not exist until it is heard in performance’ (Dusman
1994: 131). Rink himself, meanwhile, has argued the following:

Is it not the case—as I believe—that everything the performer does and thinks in the heat of C7.P4
action, coupled with what those observing and listening to the performer do and think in response
to her (or his) actions, constitute not only ‘the performance’ but also, potentially, ‘the music
itself’? In other words, when we go, say, to the Royal Festival Hall to hear Mitsuko Uchida’s
Schoenberg or Marin Alsop’s Beethoven, ‘the music’ we encounter is not limited to notes on the
page made into sounds in the air: rather, ‘the music’ is potentially de ned by our entire experience
of what is happening, encompassing everything that hits our senses.

(Rink, 2016, unpaginated: para. 3)

Finally, re ecting and to a certain extent exceeding both of these perspectives, Nicholas Cook and Richard C7.P5
Pettengill have asked: ‘after all, what is music if not performance, real-time collective practice that brings
people together as players and listeners, choreographs social relationships, and expresses or constructs
individual and group identities? […] Take away the act, take away the performance, and you take away the
music’ (2013: 1). This totalizing view of musical performance, in which music is (and only is) its own
performing, has a di erent temporal structure. Here, the musical ‘work’ (if one can still call it that) exists in
a purely local temporality: it is the sum total of all the activities taking place in the performance arena from
3
the start of the performance until the end.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
These two temporalities cut across one another in several ways. In the second (local) model, a piece is C7.P6
p. 113 entirely de ned by the events of a single performance—according to Cook and Pettengill, and Rink, that
is what the music is. In the rst (transhistorical) model, however, the piece is always in a state of becoming,
de ned by its relationship to previous performances. And these two viewpoints cannot coexist: to have a
relationship to another performance of the same work, there must already exist a work-concept—we must
know what constitutes a performance of that work, and what constitutes a not-performance of that work
(be that a performance of a di erent work, or no performance of any work whatsoever). But it is precisely
that stability that the second model seeks to unsettle. Again, this is a temporal question: how is a work to
maintain a relationship with its own past?

It is possible that the discrepancy arises merely from an evolution in Cook’s thought, and that he has C7.P7
abandoned his previous theories of a ‘horizontal’ (or, as I would have it, historical) understanding of the
musical work. Nevertheless, this evolution still betokens a wholesale move away from ontological stability
to the point where it might no longer make sense to speak of ‘pieces of music’, going further even than
considering performance as an act of creation in its own right (already a radical proposition in the long view
of musicological history), but as an event for its own sake, no matter the musical result. This, in retrospect, is
the heart of Cook and Pettengill’s strongest claim—that music is primarily a ‘collective practice’ that is
signi cant insofar as it ‘choreographs social relationships, and expresses or constructs individual and
group identities’. It may be that there are good reasons for such a radical move away from the traditional
sense of a musical work as somehow partially autonomous, partially transcendent of its own performing;
nevertheless, a theory that proposes nothing short of a revolution in the way in which the musicology of
Western art music is done merits the closest scrutiny.

Considering the problem from the angle of temporality will prove to be enlightening. So far, the work- C7.P8
concept has been closely tied, even synonymous, with the identity of the work: what in Beethoven’s
Symphony No. 9 in D minor, Op. 125 (1822–4) is ‘Beethoven’s Symphony No. 9’. I will argue that this is a
blind alley. Instead, I will attempt to separate the idea of an ontologically stable ‘core’ in a piece of music
from the di cult question of its identity, rightly and astutely problematized by Cook and his colleagues.
First, the Deleuzian colour of Cook’s description of musical performance will provide an opportunity to
explore what a Deleuzian musical work might look like: this will give rise to a case study that examines the
temporality of the work-concept in two intersecting ways. One the one hand, the deconstructive approach
taken to the musical material suggests that the work’s relationship with history is more sophisticated than
what is evoked by a simple ‘horizontal plane’. On the other hand, and triggered by Deleuze’s own invocation
of the death drive, I suggest an ontology of the musical work founded on its essential identity with human
ontology, speci cally that both are founded on a dialectical relationship with time: while ‘to be in time’ is
the precondition for existence, it is nevertheless resisted both by human self-consciousness and by the
musical artwork.

In doing all of this, I defend one of Cook’s propositions that, at rst glance, seems radically and insolubly at C7.P9
odds with the rest of the body of his theory: ‘the fantastical idea that there might be such a thing as music,
p. 114 rather than simply acts of making and receiving it, [which] is arguably the premise of the Western “art”
tradition’ (Cook 2012b: 188).
The Deleuzian Musical Work C7.S2

Cook’s understanding of the musical work has Deleuzian in ections: the rejection of the stemma, the tree, C7.P10
in favour of a network of equal relations, bears a striking resemblance to Deleuze’s rhizomatic forms.
Equally, the idea that musical performances are not performances of an original idea, but rather are
constitutive of it, recalls Deleuze’s polemic against repetition within a concept: ‘repetition is not only

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
de ned in relation to the absolute identity of the concept; it must, in a certain manner, itself represent this
identical concept. […] Repetition is not content with multiplying instances of the same concept; it puts the
concept outside itself and causes it to exist in so many instances hic et nunc’ (Deleuze 2014: 357). In other
words, Deleuze advocates the same idea of a di erence without an original. It is the ner details of Deleuze’s
approach to repetition that will make surpassing a narrow identitarian idea of the musical work possible.

Michael Gallope has already suggested what a Deleuzian work-concept might look like, situating it against C7.P11
both the traditional ‘ideal’ musical work and Goehr’s genealogy of the work-concept—which, as an
empirically veri able, historical, and regulatory framework, Gallope sees as thoroughly material. He
suggests that Deleuze’s philosophy is such a radical extension of materialism that it loops back to regain
some of the ideal qualities of Platonic idealism:

Deleuze’s philosophical orientation is so empirical, it is, in a way, purely empirical; it is so speci c C7.P12
to experience [that] it actually exalts the empirical to a new realm of purity. This has the
paradoxical result of evacuating the located speci city of experience altogether, purifying living
things down to their very materiality—a life—an unsituated, ahistorical becoming. For Deleuze,
this is the only way to think the immediacy of life to Being: by rendering life so radically
contingent that it can no longer be said to relate to any stable identity.

(Gallope 2008: 97)

In other words, the very foundation of existence is not transcendent (as Plato might have it), outside of the C7.P13
subject or object, but rather interior to it, immanent to its very nest detail. Understanding something in
purely its own terms, peering into its very materiality, reveals the ‘ground’ of being—what being ‘is’ before
it is speci ed as a person, in a place, doing something. It is looking at life in such exhaustive detail that it
becomes bare life, being such that it becomes bare being—that which is common to everything, before it
becomes di erentiated. Deleuze calls this the ‘virtual’: ‘the virtual is the presupposed ground of these
endless di erentiations. So, the virtual is really potentiality itself, the absolute potential of all life, insofar as
this creative “fuel” is nothing less than “an abstract and potential multiplicity” ’ (Deleuze 2014: 98).

p. 115 Gallope insists that recognizing this fundamental truth of existence is one of the principal functions of art: C7.P14
‘if we are going to nd anything like a musical work in Deleuze it will be one that lets us tune in to the
virtual, one that helps us escape our sedimented existence in actual, worldly relations’ (2008: 101). But,
precisely because of the nature of Deleuze’s virtual, this fundamentally changes the nature of the ideal
which is under investigation: ‘for this Deleuzian ontology of transcendental empiricism, there is no speci c
listening subject, and along with this no speci c performer, no speci c composer, and really, no speci c
musical object to speak of’ (p. 101). If a Deleuzian analysis is incapable of locating a Platonist musical work
—what might be called ‘the music itself’—what does it therefore locate? I will argue that a Deleuzian
4
perspective captures ‘music itself’, distinct from ‘the music itself’: music in general, bare music. Gallope
notes:

It is not the musical work itself that is preserved […] and it is not the composer or the performance C7.P15
or the musical culture that are preserved, it is the sonorous sensation itself that is absolutely
preserved. What is sensation for Deleuze? It is what is left of art when you subtract out all subjects,
objects, all worldly and actual attributes to art. You are left with nothing but a sensation itself—
absolute sensation.

(Gallope 2008: 104)

It is crucial to recognize that in Deleuze’s philosophy, ‘sensation’ is not understood in a vulgar aesthetic C7.P16
manner: as something heard or smelled. It is not, in other words, a retreat into the idea of music outside of
thought and language, one that reduces it to pure noise, or even pure vibrations—what might forgivably be

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
termed, that is, ‘absolute sensation’. This would necessitate a sensing subject, and a sensed object, which
have already been discarded; it would also fail to pass through and overcome the everyday materiality of the
sensed object in the way advocated by Deleuze, instead accepting that materiality—sonic vibrations,
aromatic molecules—as a given. Rather, sensation here is sensation without an object or a subject, which is
to say the very mechanics of sense: consciousness itself.

It is for this reason that I nd it di cult to agree with two of Gallope’s examples of a Deleuzian perspective C7.P17
on music. O ering a partial de nition of a Deleuzian musical work, he writes:

We can think the purity of sensation in itself as a directly un-negotiated coupling of matter and C7.P18
nervous systems. […] A Deleuzian musical work cannot be separated from the aggregate of all
nervous systems that have ever and will ever sense it. It is the self-positing, abstract and
autonomous unity of sonorous material and sensation that stands immediate to itself. (2008: 104–
105)

And illustrating what might be called a Deleuzian ‘moment’ in a musical performance, Gallope describes the C7.P19
‘spontaneous’ reaction of an audience to a semitone key change in the middle of a pop concert:

This is when chills went down the spines of the audience, bringing them into a self-organizing C7.P20
p. 116 applause machine, blowing through the airwaves like a spontaneous hurricane. This is the
moment, we might say, when the audience is not receiving meaning from a musical object or even
perceiving a musical object as such, but instead it is the moment when these millions of bodies are
a ected beyond themselves as subjects, bringing them into assembled sensational motion outside
their identity. (2008: 111)

The rst example locates the meaning of ‘sensation’ in the conjunction of sensed object (‘matter’) and C7.P21
sensing body (‘nervous system’). However, as we have seen, the Deleuzian virtual is signi cant because it is
not a conjunction of two formerly separate entities, but rather the ground which entities already share.
Thus, Gallope undervalues the achievement of the musical work in bringing the fundamental identity of
‘music itself’ and ‘sensation’ (which I align simply with consciousness) to the fore. In the second example,
however, Gallope overvalues the material surface of the music: the audience’s reaction cannot be
‘sensation’ as Deleuze would have it, not least because there can be nothing immanent about a semitone
modulation, or the response elicited by it. A semitone modulation in a pop song is highly culturally
mediated, and elicits a response not because of any intrinsic properties, but rather because of a series of
carefully negotiated musical codes.

What is required, then, is an approach that digs into the materiality of a piece of music, but in doing so C7.P22
sheds the narrow materiality of the music in its speci city, to uncover the ‘virtual’, or the ‘bare music’
itself. As Deleuze says, an artwork is properly understood as that which ‘pass[es] through the nite in order
to rediscover, to restore the in nite’ (2014: 108): this introduces another paradigm—the artwork not only
as an event happening in time, but as an act that connects with something outside of time. And it is high
time to consider a concrete example.
The Jazz Work C7.S3

At rst thought, jazz might seem the least likely place to encounter a viable work-concept. Founded as it is C7.P23
on a tradition of improvisation and quotation, of a set of standards that are passed around, played, and
recorded by dozens of di erent groups across nearly a century of songwriting, of di erent charts that share
the same harmonic structure, and of the same charts that can be played with entirely di erent

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
instrumentation, tempi, harmony, and even melody, jazz seems to shatter every casual assumption about
musical authorship and musical identity. But it is precisely for these reasons that it is the perfect landscape
in which to discover a work-concept away from the narrow concerns of identity. To understand this better,
it is useful to consider a speci c example: Misha Mullov-Abbado’s arrangement of ‘September’ (2015) for
his own quintet, covering the original by Earth, Wind & Fire (1977).

The jazz arrangement exists at some distance from the original. It begins with fragmentary punches in the C7.P24
5
p. 117 bass and piano, punctuated by interjections from the drums and wind instruments. These are revealed to
be snatches of what emerges next: a highly syncopated descending lick for the bass and piano left hand,
which will underpin most of the next section. As this lick establishes itself, the improvised punctuation from
the other instruments melts away and the drums join in the groove. So far there is nothing obviously linking
Earth, Wind & Fire’s original song, and this new jazz arrangement. Eventually, the rst shared material is
incorporated: the melody enters on the trombone and saxophone. Highly syncopated, proceeding in parallel
fths, and totally reharmonized, it is surely recognisable as the melody of the original song, but its
character has entirely changed. One could be forgiven for judging that Mullov-Abbado has not so much
arranged ‘September’, as used its raw material as the basis for an entirely di erent composition; as we shall
see, however, this would not do justice to the ethos of the piece.

An analysis of the entire track is unnecessary in this context, although it is important to draw attention to C7.P25
some salient overall features. While the ‘verse’ section always proceeds as above—all spiky syncopations
and opaque harmony—the chorus opens out into a soaring melodic alternation between the trombone and
saxophone. In a witty nod to the origins of the song, the drums momentarily provide a characteristic disco
hi-hat pattern. The harmony is more functional and the melody is delivered in a more cantabile style: that is
to say, although the texture is thinner at this point, the chorus as a whole feels more stable than what has
preceded it, as one might expect from a chorus. The same can be said of the bridge passage with its
obsessive three-note loop. In Mullov-Abbado’s version it is almost excessively bridge-like, the loop
underpinned by thick rising parallel chords, the reduction in rhythmic complexity matched by the
drummer, who o ers just the merest feathering on the cymbals. The rising harmony and continuous loop
slowly builds as the drummer thickens out the texture, cranking up the energy to send the band (and the
listener) careering into the next section.

One of the interesting things about the Mullov-Abbado, then, is the way it deconstructs Earth, Wind & Fire’s C7.P26
original, while retaining—even emphasizing—its key elements. In terms of actual musical ‘material’,
Mullov-Abbado uses precious little of the original: at most, the melody and a similar tempo. Everything else
—harmony, rhythm, structure, instrumentation, and style—is new. Given the transposition into Mullov-
Abbado’s decidedly cerebral jazz language, it is remarkable how faithful the cover seems to both the spirit
and the style of the original, and indeed it is a shock to discover, if one listens to both one after the other,
how di erent they are. It would be easy enough to assert that a careful management of texture, and e ective
quotation from the original—a disco beat here, an R’n’B lick there—is just enough to maintain a super cial
link between the two. But this would sell short both Mullov-Abbado’s accomplishment and the nature of the
work-concept I believe to be at its heart.

Considering the way Mullov-Abbado interacts with these pre-existing elements is crucial to understand my C7.P27
proposed work-concept. A clue for how to proceed is given at the very beginning of the track, when the
improvisatory punches bring ‘September’ out of mere noise: there is at once, therefore, a dialogue between
‘music’ and ‘non-music’. Alarm bells should rightly ring at this point, since the boundary between music
p. 118 and noise is a contested eld—both more widely in musicological discourse, and in this piece speci cally.
This is not a aw in the argument, however, but one of its most crucial features: none of the piece’s
referents is a discrete instantiation of stable totalities, but rather all are discursive constructions. There are
no such categories as ‘music’ and ‘non-music’ existing independently of ‘September’, and therefore there
is not a moment when the song moves from being legible within one framework to the other: the di erence

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
between them, and the movement from one to the other, is staged—and to a certain extent constructed—in
the piece.

And so it is with the piece’s other elements. ‘September’ is a disco track played by a jazz band, containing a C7.P28
wordless chorus (that is nevertheless clearly a chorus), a bridge more bridge-like than the original, and
improvised solo sections that tread a careful line between quotation and motivic development. The point is,
however, that all formal and structural categories are discursively negotiated: there is no checklist of things
that must or cannot be present to qualify as a chorus, or a disco track (or indeed a sonata exposition or
ballade—so in this, the Mullov-Abbado is not exceptional). On the other hand, by existing in such
interstitial territory, between original and borrowed, jazz and pop, and by toying with the limits of the
formal categories brought into play, this negotiated character is highlighted. It could even be said that that
very negotiation is one of the things being performed.

I argue that the continuous intimations of marginality in Mullov-Abbado’s ‘September’ lend it special C7.P29
signi cance when trying to construct a plausible work-concept. The track lies at the intersection of a
number of textual categories: disco versus jazz, improvisation versus composition, pre-existing material
versus invention, and vocal versus instrumental music. But since nearly all music will lie in the rich grey
area between these extremes, one way of reading Mullov-Abbado’s ‘September’ is as an exploitation, for
musical e ect, of the immanent fragility in all composed music. As Cook and his colleagues are keen to
articulate, no music is coterminous with its score, since convention and improvisation play a part in all
musical performance; more than this, Barthes reminds us that all text is simply a ‘tissue of quotations’
(Barthes 1977: 146). Music is no exception: with only twelve notes to choose from, already layered in a para-
tonal hierarchy, much of music’s supposedly immanent meaning is merely the transmission of socially
determined factors, which composers can manipulate and sometimes undermine, but never truly escape.
And so the easy argument against Cook’s ‘historical’ conception of the work in Beethoven’s Symphony No.
9 is turned on its head: his formulation is problematic not because it is too broad, but because it is too
narrow. A contemporary performance of Symphony No. 9 does not only refer to its previous performances,
recordings, and scores, but to all music, the entire tonal tradition, and indeed the entire edi ce of art music
itself—including the socially determined de nition of what constitutes ‘art music’. This is an impossibly
large eld to consider: it is coterminous with language itself, the very condition of possibility of human
thought, and therefore must be rejected. Instead, it is necessary to focus on the way music pulls these
frames of reference together, and this is where Mullov-Abbado is particularly useful.

p. 119 My invocation of language is not accidental: for all the freedom on display during the Misha Mullov-Abbado C7.P30
Group’s performance, the overriding sense one receives as an audience member is of an e ort to
6
communicate (or as Ingrid Monson might put it, to ‘say something’). A particular understanding of
language, as the very condition of possibility of being-in-the-world and the only means of materializing
inarticulate Being, is crucial to my argument, and I contend that Mullov-Abbado and his group stage this
coming into being at various levels in their performance of ‘September’. Most obvious is the gradual
emergence of the groove from the track’s primordial beginnings, pulling composed music out of semi-
random noise, that literally performs the act of making music. From here, the move from Mullov-Abbado’s
entirely original bass lick into his re-imagining of Earth, Wind & Fire does not so much push the song into
an additional discursive arena (the arena of music history) as drag that entire discursive arena into the
performance—into being.

This deconstructive angle obviously supplants any Platonic idea of the musical work: it does not make sense C7.P31
to posit the existence of some ideal version of Mullov-Abbado’s ‘September’ for which the performances
and recordings are material evidence, since it is this very movement from ideality to materiality that is
crucial to the reading above; clearly, it makes even less sense to think of Mullov-Abbado’s version as an

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
instantiation of Earth, Wind & Fire’s ur-September. On the other hand, a Cookian historical understanding
of the musical work does not seem appropriate either. While a comparison of several performances of
Mullov-Abbado’s ‘September’ over time may point to an of orbit of variants within which the piece’s
identity is the absent centre—a ‘work-concept’ in a banal sense—given the tantalizing possibilities
glimpsed above of exploring the piece’s relationship, not only with its own past but with the very concept of
‘the past’, pressing on with a deconstruction of musical identity seems more worthwhile.

The discursive arena I have alluded to—that which puts the musical material in the past tense, or at least in C7.P32
quotation marks—comprises both the ‘borrowed’ content of Earth, Wind & Fire and formal categories like
7
‘chorus’ and ‘bridge’. A subtle play between signi cation and its own undermining is at work in the formal
structure of the song and in its content, suggesting both that music is related to its past at a more
fundamental level than simply the identity of its raw material (complicating the distinction within a piece
between music’s present identity and its past) and that this very relationship is discursively negotiated in
performance (blurring that easy distinction, by pushing the past into the present).

For example, Mullov-Abbado’s wordless chorus does not put into action a selection of techniques against a C7.P33
background of predetermined chorus possibilities, nor does it detract from them, either by omitting
putative ‘chorus-like’ tropes or introducing ‘non-chorus’ features. Are harmonic stability and cantabile
delivery invariably chorus tropes—that is, if the dissonance of the outer sections continued, and were the
tune given to the piano in spiky block chords, would it cease to be a chorus? No: the deployment of these
musical features in the context of the rest of the song circumscribes the way in which this section is a
chorus, they do not de ne what one is. Even those features that may seem counterintuitive for a chorus
p. 120 section—textural reduction, and of course the absence of words acting as a repeated refrain—do not
undermine its status as a chorus, as much as they undermine the assumptions of what a chorus is.

Likewise, there is no de nitive record of what elements are proprietary to Earth, Wind & Fire, no stable past C7.P34
to draw upon like an archive, and therefore no sense of where Earth, Wind & Fire ends and Mullov-Abbado
begins. The hi-hat lift that Mullov-Abbado so wittily deploys to refer to the song’s disco history is not
exclusively a disco technique, nor is it even essential in the disco genre: indeed, rather ironically, there are
no hi-hat lifts in the original recording’s drumbeat, in which the hi-hat is only played closed. Nevertheless,
as a semiotic clue it clearly references the song’s immediate musical past by invoking a speci c style: but it
is now obvious that this past is a construction, or narrative, rather than an artefact or relic. In this way,
Mullov-Abbado de nes not only his relationship to that past but that past’s relationship to itself. The
Deleuzian echoes are clear: the musical material is not reliant on an ideal sca old or supplement to give it
meaning, and the musical cues do not point exclusively outside themselves to a stable backdrop from which
the listener draws interpretative sustenance. Rather, the very peculiarity of the musical material itself, once
we dig into it, gathers and structures its own context, its own concept. And in the process, a linear concept
of music-historical time extending back into the past and forward into the future has been supplanted by an
entirely di erent model, one based solely on the turning point between when a piece does not exist, and
when it does—when it is ‘in time’, and when it is not.
To Be in Time C7.S4

I suggested above that a Deleuzian understanding of repetition would be helpful in developing the C7.P35
contemporary work-concept, since it is founded on the same ‘horizontal’ principles encountered in the
study of musical performance: ‘repetition without an original’. This still leaves a question—what is
repeating? What is insistently pushing itself into time? Deleuze argues that the ‘secret verticality’, the true

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
source of repetition, is ‘furnished by the death instinct’. This is the keystone to a properly Deleuzian
understanding of the musical work, and requires detailed explication, since it requires us to tie any ontology
of music to a broader human ontology. The ‘death instinct’ (or ‘death drive’, ‘Thanatos’) is, following
Freud, the fundamental ground of human consciousness: the tendency of human beings to die (see Freud
2015[1920]). It is not necessary to enter into debate about the precise nature of the death drive and its
opposite, the pleasure principle (‘Eros’), beyond noting that their duality persuasively stages a truth of the
human condition: that consciousness is a gross excrescence during what is otherwise a simple chemical
reaction of growth, reproduction, and decay common to all other processes in the observable universe, and
thus both its arbitrariness and its futility are de ning features. Living, thinking, existing as we understand
it, are forces of sheer will: an ongoing process of self-creation—a constant ‘repetition’—of an organism
thrust unwillingly into consciousness, which is to say an awareness of time. Heidegger put it
p. 121 uncharacteristically concisely when he said that a human—in his terms, Dasein—is ‘distinguished by the
fact that, in its very Being, that Being is an issue for it’ (Heidegger 1962: 32). In other words, humans are
conscious of their own Being: we are aware we exist.

It is a commonplace that music is one of the ways humans narrate their own existence, literally giving C7.P36
signi cance to the passing of time, by making time itself signify; as Susanne Langer put it, ‘music makes
time audible, and its form and continuity sensible’ (Langer 1953: 110). Along these lines, therefore, it would
be a straightforward claim that the ultimate arbitrariness of the tonal or chromatic system mirrors the
arbitrariness of conscious existence, as does the drive nevertheless to imbue both with meaning. J. P. E.
Harper-Scott has criticized this super cial understanding of Heideggerian ontology (which I argue, for the
purposes of this chapter, is of a piece with Deleuze’s): ‘poststructuralists would agree with Heidegger’s
argument that, as a language-bearing being, man invests the world with meaning, but they run to a
di erent conclusion. For the poststructuralist, it therefore follows that the world is either intrinsically
meaningless, or else that such meaning as it has ought not to be trusted, because it was made up by a
controlling elite’ (Harper-Scott 2006: 166). In a properly Heideggerian reading, however, precisely what is
‘human’ about humans is understanding. Meaning may be incomplete, but it is never absent: not only is it
all we have, it is all we are: ‘on [Heidegger’s] view, we are always understanding, never not understanding’
(Harper-Scott 2006: 156). Or as Deleuze says, more simply, ‘the Self itself is a contemplation’ (Deleuze
2014: 97).

Thus, the relationship between the arbitrariness of existence (signi ed by death) and the search for C7.P37
meaning (signi ed by humanity’s capacity for self-narration) are not so much contradictory as they are
dialectically paired. The drive towards what we call death is positive, and necessary: it is a chemical reaction
just like the burning of a star, or acid rain. During that chemical process, however, consciousness arises,
something which is experienced as resisting the forward ow towards completion—or death. It is only
consciousness which stages (the beginning of) the end of that reaction negatively as ‘death’, but it is only
that reaction itself—which must end in ‘death’—that makes consciousness possible. Living is experienced
as ‘Eros’, a forceful repetition—repetition being the opposite of development, progress (or ‘decay’)—but it
takes place over a ground of ‘Thanatos’, the ineluctable rush towards ‘death’. In Deleuze’s more poetic
language: ‘Eros and Thanatos are distinguished in that Eros must be repeated, can be lived only through
repetition, whereas Thanatos (as transcendental principle) is that which gives repetition to Eros, that which
submits Eros to repetition’ (2014: 22).
This immediately has consequences for a coherent theory of temporality: the focus on repetition, on C7.P38
continuous re-creation, forces everything into the present tense: ‘the synthesis of time constitutes the
present in time. It is not that the present is a dimension of time: the present alone exists. Rather, synthesis
constitutes time as a living present, and the past and the future as dimensions of this present’ (2014: 101).
That is to say, Deleuze’s ontological argument presupposes the deconstructive-discursive one above:
understanding humanity as a consciousness embodied in language, the past is not an archive to which one
contributes or a background against which one performs, but rather something that is drawn up and pulled

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
p. 122 into time during those very acts of repetition. Crucially, the comparison with music exceeds mere
analogy. As a locus of human ‘contemplation’—a speech act of sorts, that structures the present by
structuring the past—musical performances are not metaphors for human living, they are a primary
instance of it.

Therefore, what is special about music is that it performs precisely what is commonplace: the repetition of C7.P39
existing. Deleuze and Guattari say that ‘if successful, [the] monumental and permanent work produces a
metaphysical incarnation of what all music can do at any point in space or time, that is, produce an absolute
“sensation in itself”, stripped of all worldly mediation and interpretation, revealing the ontological variety
of all there is’ (Gallope 2010: 111). This problematic word, ‘sensation’, can now be properly understood: the
Self as contemplation, Dasein, consciousness. What is special about the ‘sensation’ that musical works bring
about is, paradoxically, not musical at all, but absolutely general: it is that fundamental act of literally
bringing an object of contemplation into Being (i.e. into the present, into time). Moreover, to be ‘stripped of
all worldly mediation and interpretation’ does not mean that music is experienced in an ahistorical,
undi erentiated void, but the exact opposite: as I have shown, it is the very speci city of the music’s
materials, and the means of their deployment, that point up their relationship with the rest of history—with
‘the ontological variety of all there is’.

Beyond Time: Towards ʻImmortalityʼ (by Way of Conclusion) C7.S5

This chapter began by considering the question of musical identity: what makes Beethoven’s Symphony No. C7.P40
9 ‘Symphony No. 9’. In place of this question, a theory of musical ontology has emerged that aligns it with
human ontology. I have argued that it is more illuminating to consider musical performances as acts of
contemplation, and as such inherently temporal: contemplation here is understood as a continuous act of
‘repetition’, contextualizing the present by gathering and structuring the past and future, in an attempt to
rationalize the otherwise purely chemical process of not dying. Nevertheless, to tie up all the loose ends it
still remains to zero in both on what (or where) the musical ‘work’ is and on the point at which the speci cs
of musical ‘identity’ can usefully be jettisoned.

Hannah Arendt’s nuanced distinction between ‘labour’ and ‘work’ is useful in this undertaking. For Arendt, C7.P41
‘labour’ is that which is required simply to maintain life: those things that are immediately used up, and
have no permanence—food is eaten, clothes wear out. ‘Work’, however, is that which makes merely existing
into living: ‘viewed as part of the world, the products of work—and not the products of labour—guarantee
the permanence and durability without which a world would not be possible at all’ (Arendt 1958: 94). Arendt
recalls the Ancient Greek distinction between zoē and bios, ‘bare life’ and ‘living’: labour sustains zoē, the
p. 123 chemical act of not dying referred to above, whereas work relates to bios, which following the Deleuzian
discussion above we might also term Eros, or repetition. ‘The chief characteristic of this speci cally human
life […] is that it is itself always full of events which ultimately be told as a story, establish a biography; it is
of this life, bios [living] as distinguished from mere zoē [life], that Aristotle said that it “somehow is a kind
of praxis” ’ (Arendt 1958: 97). The resonances with the ontology elaborated above, which constructs human
life as a dialectical pairing of biological ground and a fundamentally linguistic metanarrative, are obvious.
But in her provocative use of the word ‘permanence’, Arendt invites us to go even further in considering the
temporal quality of human (and musical) ontology.

Arendt is clear that humanity’s ideal existence is entirely separable from its animal foundation: ‘we need C7.P42
not choose here between Plato and Protagoras, or decide whether man or a god should be the measure of all
things; what is certain is that the measure can be neither the driving necessity of biological life and labour
nor the utilitarian instrumentalism of fabrication and usage’ (1958: 174). She is equally clear that the art

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
work emanates from this ideal realm—in a way directly comparable with Deleuzian sensation:

The immediate source of the art work is the human capacity for thought […]. These are capacities C7.P43
of man and not mere attributes of the human animal like feelings, wants, and needs, to which they
are related and which often constitute their content. Such human properties are as unrelated to the
world which man creates as his home on earth as the corresponding properties of other animal
species.

(Arendt 1958: 167)

Thus it is clear that the art work is permanent, to use her own terminology, because it emerges from a C7.P44
peculiarly timeless part of human life: ‘it is as though worldly stability had become transparent in the
permanence of art, so that a premonition of immortality, not the immortality of the soul or of life but of
8
something immortal achieved by mortal hands, had become tangibly present’ (p. 168). In this way, both
humans and their artworks can be considered ‘permanent’, not because they never decay, but because their
decay is irrelevant to them. Both are de ned by the opposing pole of the dialectical pair, the one that
persists despite death and decay: repetition. What is human about humanity is sensation, above and beyond
animality; what is ontologically stable about music—what is workly—is that which aligns it with human
thought and human language.

This, it seems, is what Cook means when he talks of ‘the fantastical idea that there might be such a thing as C7.P45
music, rather than simply acts of making and receiving it, [that] is arguably the premise of the Western
“art” tradition’ (Cook 2012b: 188). It does, however, run counter to the overall thrust of his
conceptualization of the musical work, for, if this is the case, then when music is ‘choreographing social
relationships’, or ‘expressing […] group identities’, then it is not being a musical work in this new sense.
Under this new de nition, music is a ‘work’ insofar as it is decidedly more than its performing, more even
than its identity. In a true Deleuzian sense, the material level has been penetrated so fully that it is ‘puri ed’
of its speci city, reaching a form of ideality. The question of musical identity has been entirely (but I believe
p. 124 productively) begged: the musical work is not the bit of Beethoven’s Symphony No. 9, ‘Symphony No. 9’,
neither is it a discursive accumulation of a history of performances, nor all its sounding, moving, and
hearing parts. The consideration of Mullov-Abbado’s ‘September’ concluded with the distinction between
when a piece of music exists and when it does not—the act of throwing a work into the world, of gathering
the world around a musical act, of bringing something into time. It is precisely in this movement that it is
possible to maintain some of the discrepancies in Cook’s various formulations of the musical work—
especially that between music (only) as performance, and as a transcendent work—not as irresolvable
con icts but dialectical pairs. The musical work is an act, just as Arendt reminds us (via Aristotle) that living
is a kind of praxis. But the musical work is ‘permanent’, and humans are ‘immortal’, because these acts
point outside of themselves; indeed, one of their functions is to intimate a world beyond acts, a world which
makes acts possible in the rst place.

And therefore music’s workly quality is in fact the element that is never performed, that which is always left C7.P46
behind. It is that part that could never be performed, because it exists apart from itself, emerging from bios
—that ontological plane that, for humans to be in time, must remain outside of it.
Notes
1. Alongside the works cited throughout this chapter, Cookʼs conclusions on musical performance are gathered in his C7.N1
monograph study, Beyond the Score: Music as Performance (Cook 2013).

2. For prescriptive vs descriptive types of performance studies, as well as a useful overview of the whole subdiscipline, see C7.N2
Rink (2002: esp. 35–39). Among the ʻfoundationalʼ texts that he draws upon are Berry (1989), Dunsby (1995), and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
Schmalfeldt (1985).

3. Here it is already possible to glimpse one of the fundamental obstacles to such a conceptualization of the music ʻitselfʼ, or C7.N3
the musical ʻworkʼ: denuding the music of all objecthood, and defining performance in purely temporal terms, makes it
almost impossible to draw boundaries around it. What is the ʻperformance arenaʼ—how far does it extend? That is to say,
where does the performance stop—at the edge of the stage, or the auditorium, or even the concert hall bar, where the
music may still be audible through the walls or over a tannoy? Even the temporal borders are di icult to police: does the
piece end at the double barline, with the applause, the discussion in the pub a erwards, or as the audience members
relive the music lying in bed later that night? This radically temporal perspective has no intuitive cut-o point, which—if
taken to its logical conclusion—threatens to entirely destabilize the identity of pieces of music themselves. I will go on to
argue that only by begging the question of ʻidentityʼ at an even more profound level will it be possible to reconcile the
several competing conceptualizations of music outlined in the first part of this chapter.

4. It is important to note that ʻbare lifeʼ without specific subjects, or ʻbare musicʼ without any historical specificity, is not C7.N4
some sort of pre-social Nirvanic proto-consciousness. Rather, it is Being (capitalized, as in Heidegger) entirely in excess of
the social: it is the very ground of sociality, insofar as it constitutes a commonality of Being shared by all people, before
they become fractured by the di erences that constitute society.

p. 125 5. In live performances the opening can be even freer than the album take, consisting of free improvisation and C7.N5
incorporating extended techniques and noise elements, before composed elements are gradually introduced.

6. Monson (1996). This chapter is at some remove from the ethnographic approach at work in her study, but of course in C7.N6
total agreement with her underlying premise that musical performance puts into practice fundamental truths about the
nature of human societies, and as such is worthy of the greatest scrutiny, even if the discursive richness is not a primary
concern of—or even remotely interesting to—the performers and audiences themselves. An approach that is much closer
to the perspective of this chapter, however, is Okiji (2018), particularly in its examination of how jazz contests notions of
individuality, and in its critical re-reading of ideas of surface and depth in the jazz work (see especially chapter 1). Equally
compelling is Okiji's reading of jazz as a music that is not coincident with the world, but rather tugs away from it,
highlighting its partiality (chapter 3); it is, however, outside the scope of this author to develop—as Okiji does—the way
this intersects with black, especially Black American, experience.

7. This is a list which could be continued ad infinitum: ʻsongʼ, ʻtonalityʼ, ʻjazzʼ, ʻmusicʼ…since all are historically mediated. But C7.N7
one of the aims of this chapter is to show that, rather than extending endlessly and unpredictably, there exists a vanishing
point from which all of these historically mediated groups emanate, located in a Deleuzian concept of sensation: the
ʻworkʼ.

8. The idea of humanityʼs permanence has since been taken up in more recent philosophy, specifically that of Alain Badiou C7.N8
(see Badiou 2012: esp. 10–13). Arguing against a conception of humanity that ʻequates man with his animal substructureʼ
and ʻreduces him to the level of a living organism pure and simpleʼ (p. 12), he has proposed instead that ʻMan is to be
identified by his a irmative thought, by the singular truths of which he is capable, by the Immortal which makes him the
most resilient and most paradoxical of animalsʼ (p. 16). Again, this is founded on the fundamental duality of human
ontology that Badiou, too, perceives: ʻaccording to human finitude, two situations are separable: those which are
subsumed under the attribute of thought (cogitatio) and those under the attribute of extension (extensio). The being of
this particular mode that is a human animal is to co-belong to these two situationsʼ (Badiou 2013: 119).
References C7.S6

Arendt, H. (1958). The human condition. University of Chicago Press. C7.P47


Google Scholar Google Preview WorldCat COPAC

Badiou, A. (2012). Ethics: An essay on the understanding of evil (trans. P. Hallward). Verso. C7.P48
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
Badiou, A. (2013). Being and event (trans. O. Feltham). Bloomsbury. C7.P49
Google Scholar Google Preview WorldCat COPAC

Barthes, R. (1977). The death of the author. In S. Heath (ed. and trans.), Image, music, text. Fontana. C7.P50
Google Scholar Google Preview WorldCat COPAC

Berry, W. (1989). Musical structure and performance. Yale University Press. C7.P51
Google Scholar Google Preview WorldCat COPAC

Cook, N. (2012a). Introduction: refocusing theory. Music Theory Online 18(1). C7.P52
https://mtosmt.org/issues/mto.12.18.1/mto.12.18.1.cook.html.
Google Scholar WorldCat

Cook, N. (2012b). Music as performance. In M. Clayton, T. Herbert, and R. Middleton (eds), The cultural study of music: A critical C7.P53
introduction, 184–194. Routledge.
Google Scholar Google Preview WorldCat COPAC

Cook, N. (2013). Beyond the score: Music as performance. Oxford: Oxford University Press. C7.P54
Google Scholar Google Preview WorldCat COPAC

p. 126 Cook, N., and Pettengill, R. (2013). Introduction. In N. Cook and R. Pettengill (eds), Taking it to the bridge: Music as performance, C7.P55
1–19. University of Michigan Press.
Google Scholar Google Preview WorldCat COPAC

Deleuze, G. (2014). Di erence and repetition (trans. P. Patton). Bloomsbury. C7.P56


Google Scholar Google Preview WorldCat COPAC

Dunsby, J. (1995). Performing music: Shared concerns. Clarendon Press. C7.P57


Google Scholar Google Preview WorldCat COPAC

Dusman, L. (1994). Unheard of: Music as performance and the reception of the new. Perspectives of New Music 32(2): 130–146. C7.P58
Google Scholar WorldCat

Earth, Wind & Fire (1977). September [vinyl 7-in. single]. Columbia. C7.P59

Freud, S. (2015) [1920]. Beyond the pleasure principle. Dover. C7.P60


Google Scholar Google Preview WorldCat COPAC

Gallope, M. (2008). Is there a Deleuzian musical work? Perspectives of New Music 4(2): 93–129. C7.P61
Google Scholar WorldCat

Gallope, M. (2010). Cavell and Deleuze. Journal of Music Theory 54(1): 107–120. C7.P62
Google Scholar WorldCat

Harper-Scott, J. P. E. (2006). Edward Elgar: Modernist. Cambridge University Press. C7.P63


Google Scholar Google Preview WorldCat COPAC
Heidegger, M. (1962). Being and time (trans. J. Macquarrie and E. Robinson). Blackwell. C7.P64
Google Scholar Google Preview WorldCat COPAC

Langer, S. (1953). Feeling and form: A theory of art. Scribnerʼs. C7.P65


Google Scholar Google Preview WorldCat COPAC

Monson, I. (1996). Saying something: Jazz improvisation and interaction. University of Chicago Press. C7.P66
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469335 by National Science & Technology Library user on 26 May 2023
Mullov-Abbado, M. (2015). September. On New Ansonia [CD]. Edition. C7.P67

Okiji, F. (2018). Jazz as critique: Adorno and Black Expression Revisited. Stanford University Press.
Google Scholar Google Preview WorldCat COPAC

Rink, J. (2002). Analysis and (or?) performance. In J. Rink (ed.), Musical performance: A guide to understanding, 35–58. Cambridge C7.P68
University Press.
Google Scholar Google Preview WorldCat COPAC

Rink, J. (2016). Response. Music Theory Online 22(2). https://mtosmt.org/issues/mto.16.22.2/mto.16.22.2.rink.html C7.P69


Google Scholar WorldCat

Schmalfeldt, J. (1985). On the relation of analysis to performance: Beethovenʼs Bagatelles Op. 126, Nos. 2 and 5. Journal of Music C7.P70
Theory 29 (Spring): 1–31.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
CHAPTER

8 Distracted Attention, Temporal


C8 Switches, and the
Consolations of Performing 
Anthony Gritten

https://doi.org/10.1093/oxfordhb/9780190947279.013.4 Pages 127–C8.P120


Published: 08 December 2021

Abstract
Distraction is frequently blamed for interfering with the ergonomic production of capital, for
encouraging substandard performance. Indeed, it is frequently con gured as an impediment to
timekeeping, a thorn in the side of consciousness, a drag on intentional action, and a brake on
decision-making. Reality, however, is complex. While distraction can interfere with timing, anxiety,
memory, error, and fatigue, it can also be exploited under controlled conditions to enhance
performance by helping the performer to maintain an open cognitive and physical responsiveness to
the world and a pragmatic mode of engagement with the task at hand. Indeed, distraction ensures that
the performer is in close contact cognitively and socially with the full phenomenological plenitude of
sound, thereby contributing to performance’s transformative value as a way of accumulating social
capital in everyday life.

Keywords: attention, concentration, distraction, performativity, performance, temporality


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

I can do many things at once: stand in line, / C8.P1

listen to the music, have ideas, wait for the / C8.P2

next conversation. C8.P3

John Cage C8.P4


Unconsoled Performers C8.S1

I open with a passage from Kazuo Ishiguro’s novel The Unconsoled. In this passage the central character, C8.P5
Ryder, who is a famous concert pianist, has been nding it hard to concentrate on preparing for a
forthcoming recital somewhere in central Europe. At one point he re ects as follows (Ishiguro 1995: 328):

Parkhurst continued in this vein for a while longer, but I had stopped listening. For his mention of C8.P6

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
my ‘complacency’ had triggered something, causing me suddenly to remember that my parents
were due shortly to arrive in the city. And there came over me, there in Miss Collins’s front parlour,
seizing me with an icy panic that was almost tangible, the realisation that I had not prepared at all
the piece I was to perform before them this evening. Indeed, it was several days, perhaps even
p. 128 weeks, since I had last touched a piano. Now here I was, only hours from this most important of
performances, not even having made arrangements to rehearse. The more I thought over my
situation, the more alarming it appeared. I saw I had allowed myself to become far too preoccupied
with the talk I was to deliver, and somehow, unaccountably, had neglected the more fundamental
matter of the performance. In fact I could not for a moment even remember which piece I had
decided to play. Was it Yamanaka’s Globe-structures: Option II? Or was it Mullery’s Asbestos and
Fibre? Both pieces, when I came to think about them, were disturbingly hazy in my mind. Each I
remembered, contained sections of great complexity, but when I tried to think further about these
passages, I found I could recall almost nothing. And meanwhile, for all I knew, my parents were
already here in the city. I saw there was not a minute to be lost, that whatever the other calls on my
time I had rst to secure for myself at least two hours of quiet and privacy with a good piano.

There are many other such passages in the novel (e.g. p. 426). In order to avoid getting distracted by literary C8.P7
and novelistic matters or the portrayal of high modernism, I claim that this situation is paradigmatic, not
exceptional, for what it narrates about thought, and musical thought in particular. While Ishiguro’s central
plot contrivance—that Ryder nds it increasingly hard to remember whom he promised to meet and where
he should meet them and he becomes increasingly unable to manage his a airs both musical and social—
plays over more than 500 pages, my reason for invoking Ryder’s predicament with respect to himself is to
highlight the fact that he is not alone in nding it hard to focus on what must be done. I seek in this chapter
to understand what is happening to Ryder and the many millions of other citizens in the developed world
whose lives seem beyond their full control (even for the best multitaskers like Cage), who nd themselves
increasingly reactive to events (their subjectivity having become not just distributed but a function of
something else, no longer the determining factor), and who nd that their thoughts are forever being pulled
in multiple directions. This includes musicians.

What creates this situation? Distraction. This phenomenon has become increasingly pervasive with the rise C8.P8
of cognitive capitalism, the attention economy, the digital turn, the increase in the planet’s human
population and their gradual convergence on cities and megalopoles. The seemingly inexorable ‘becoming-
musical’ of the world—the stranglehold that sounds have over our daily lives—is a central contributing
factor to distraction’s ubiquity. It is no coincidence that a recent philosophical theory of sounds de ned
sound itself in ways that overlap signi cantly with how most studies have approached distraction: sounds
are ‘events in which a moving object disturbs a surrounding medium and sets it moving’ (O’Callaghan 2007:
61). The extraordinary frequency and extent to which distraction impacts on everybody both individually
and collectively is shown by estimates of its nancial impact: one estimate of the cost of ‘interruptions’ to
the US economy is in the region of $588 billion per year (Trafton and Monk 2007: 111). Another researcher
admits,

Although there are certainly bene ts to having access to the rich landscape of spontaneous C8.P9
thoughts for the purpose of creative incubation, problem solving, and goal setting, an inability to
p. 129 focus attention in the face of irrelevant distraction by such thoughts can be problematic.
Unfortunately, humans have been shown to experience this intrinsic undercurrent of spontaneous,
self-generated thought during ongoing task demands as a form of interference, distraction, or
rumination approximately 50% of each waking day.

(Vago and Zeidan 2016: 97)

In short, we can debate whether ‘Man’s almost in nite appetite for distractions’ (Huxley 1958: ch. 4) works C8.P10

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
for human bene t or against it (evidence for both sides is noted below), and whether I even discuss the same
thing when juxtaposing economic ‘interruptions’ and Huxley’s view of cultural ‘distraction’, but there is no
undoing The Distraction Turn.

After all, distraction is as ubiquitous as the performance events into which it intrudes. Consider an example. C8.P11
At the end of his recital at the Wigmore Hall in London on Friday 13 November 2009, Marc-André Hamelin’s
second encore was Haydn’s Fantasia in C major Hob XVII4 Op. 58 (1789). When Hamelin arrived at the rst
of several general pauses on low unison octaves, there was a short burst of music somewhere outside the
building in a broadly consonant key area. Hamelin smiled, raised his head to listen, and the audience
giggled. At the next general pause Hamelin waited on the low note, perhaps a little longer than necessary, as
if to signal that he was waiting for the interruption, but the noise failed to occur. The audience laughed
anyway, right on cue. This example shows how distraction is both embedded into musical texts as a
possibility and yet is always a deus ex machina outside the performer’s conscious control. It is a good
example of how playing on stage does not preclude listening to noises outside the hall. How can we
understand that the constitution of this event is centred on the fact that something happened unbidden,
unplanned, unbeknownst to the performer? A summary of the standard answer to this question is given in
Roberts’s book on Debussy’s piano music, where it is claimed (1996: 286): ‘Without concentrated listening,
the ngers, arms, back, and feet may as well be blocks of wood.’ As sensible as this seems—and the threat of
being wooden when performing drives many pedagogical programmes—much of signi cance remains to be
said about ‘concentrated listening’, about its relation to distraction, and about whether the performer is as
true to her word on stage as she claims in the teaching studio and practice room.

A note on terminology. I stick with the term ‘distraction’ in preference to cognate terms used in the C8.P12
literature, like ‘interruption’, ‘intrusion’, and ‘intervention’. And by ‘distraction’ I usually mean events of
auditory distraction, when not referring to spectacles of cultural distraction. My reason for sticking with
‘distraction’ is simple, although it complicates the plot: the term embraces a wider variety of connotations
and embodied meanings across the multiple discourses that I invoke below. Sticking with ‘distraction’
a ords me the opportunity to interweave cognitive, cultural, and musical issues together in a single thread.
Thus, the term as used below means several things according to the context, such as when I cite research on
‘interruption’. In this way the argument may seem to get distracted from itself as it develops, but if this
happens then it will be appropriately thematic. Furthermore, it is clear below that the various terms used
p. 130 across di erent discourses mean di erent things. My position is that ‘distraction’ is a master term that
subsumes interruptions and other phenomena, since the latter are context-speci c and predicated upon an
economic model of cognition, or at least an economic telos for task performance—whereas I seek to unpack
the phenomenology of distraction: what it feels like for the performer to be distracted.
Ubiquitous Events C8.S2

Distraction is ubiquitous. All human systems are riven with distraction. Everybody is a ected by distraction C8.P13
at some point in one form or other, including scholars (Dell’Antonio 2004; Szendy 2008). Even Kant
complains about being ‘distracted by illness’ and describes the phenomenon of distraction in his
Anthropology (Lyotard 1989: 324). An indication of the scale of distraction can be given by listing synonyms

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
and noting their interrelationships: aberration, abstraction, agitation, amusement, beguilement,
bewilderment, commotion, complication, confusion, disorder, dispersion, disruption, dissipation,
disturbance, diversion, divertissement, engrossment, entertainment, frenzy, game, interference,
interruption, pastime, perplexity, preoccupation, recreation, seduction. There are also cognate concepts like
rapture, trance (Herbert 2011a), and mind wandering. All this is without adopting North’s (2011) more
productive approach in The Problem of Distraction (more about this later). This wide-ranging constellation of
terms could be used to sketch a detailed typology of distraction. Trahir is a French verb meaning variously to
betray, to give away, to reveal, which suggests something about the nature of distraction (these meanings
will return when I describe distraction in terms of a ‘switch’), and in particular why it is often evaluated in
moralistic terms, as with the French Moralists, who blame many social faults on the human susceptibility to
distraction and its common consequences (e.g. La Rochefoucauld 1959: §§154, 215, 219, 243, 266, 267, 398,
482, 487, 512, 630).

Distraction has a chequered history: absent from some discourses, present in others. It is generally absent C8.P14
from classical phenomenology, with the possible exception of Gurwitsch. Husserl generally does not
consider it worth his attention, even though his theory of time consciousness is predicated on precisely an
Augustinian position vis-à-vis the fragility of human attention. It might be expected that Nancy’s Listening
(2007) would be promising in this regard, but it does not really engage with distraction, even though his
con guration of ‘resonance’ would have served as a provocative starting point. Distraction has a long
history in psychology, particularly in Freud: recall concepts like ‘inattention’ and the ‘wandering’ mind
(Marder 2009: 43–45). It has a signi cant presence in cognitive science, where studies of perception have
much to say about it, often via cognate terms like ‘interruption’ and ‘interference’ (Beaman and Holt 2007;
Lavie et al. 2004; Ponjavic-Conte et al. 2013; Liebl et al. 2012). It is also discussed in certain areas of post-
Heideggerian phenomenology, especially the recent interface between phenomenology and neuroscience.
Obviously in engineering and the applied sciences distraction is a crucial touchstone against which
p. 131 applications are tested. In the humanities, distraction is discussed with respect to literature and reading,
all the way back to the eighteenth-century penchant for ‘diversions’ (Phillips 2016; Wood 2011). There is a
fair amount of work in theatre studies, notably the work of Home-Cook (2015), and the rise of interest in
‘atmospheres’ may generate further work on the subject. There are many studies about distraction in lm
that follow in the footsteps of Kracauer (1995) and, in the broader discourse of critical theory, Benjamin
(2008) and Adorno (1991) (who evaluate the redemptive potential of distraction quite di erently).

In music performance studies, however, there is no equivalent level of discussion of distraction, except— C8.P15
and often by proxy—as a feature of the cultural spectacles studied extensively in broadly
ethnomusicological (Nettl 2005) and sociological discourses (which are not my focus) (Crawford 2009).
Even a seminal text like McKenzie’s Perform or Else: From Discipline to Performance avoids explicit discussion
of distraction, and cognate terms like ‘interruption’, ‘interference’ and ‘disruption’ make sporadic
appearances (2001: respectively 208, 261, 265, 277, 204; 53, 60, 230, 266), as if e ciency can only be
measured as a function of success and not as a function of distraction, despite the fact that the central
concept of performativity is precisely a matter of how performing engages with the material world and
negotiates its force as an utterance. What discussion lacks is less the notion of music ‘as’ a distraction, and
more the distraction ‘of’ music (in the genitive case). Even in the discourse on entrainment, where
discussion is of a process that accommodates and adjusts to rhythmic complexity (Clayton et al. 2005), the
desire is often for greater and tighter entrainment. It is curious that distraction is deplored in music
discourse, given that it plays a vital role elsewhere in contemporary life, as with warnings and alarms,
especially since hearing is sometimes called ‘the sentinel of the senses.’ (Banbury et al. 2001: 12–13)

Indeed, distraction is by and large absent from the discourse of performing music, as if there is an unwritten C8.P16
rule that its very name must never be uttered. This is curious, because (as is clear from what follows),
although the preferred term used to describe the performer’s listening is ‘attention’ or a cognate like

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
‘concentration’, ‘absorption’, or ‘focus’, attention is not the opposite of distraction, as Cage concluded
repeatedly. Con guring performing in terms of attention is not the opposite of con guring it in terms of
distraction. There are four reasons for the relative downplaying of distraction in studies of performing. (1)
The rst reason is the assumption that distraction is marginal by comparison with attention, concentration,
and absorption. Thus, in terms of McKenzie’s (meta-)theory of performance, ‘perfumance’ (2001: 193–275,
esp. 228–236), even the central concept of ‘performativity’ ends up being constrained to being itself
performative, i.e. attentive to the performance task, rather than distracted and genuinely open to what
happens. (2) The second reason is that performing music (the other performing arts having long since
expanded their conceptual position) is still very much bound up with the logic of ‘performance of a text’, as
opposed to ‘performance from a script’, numerous deconstructions notwithstanding. There are
longstanding cultural reasons for this, such as the institutional set-ups of music conservatoires, which
generally do not encourage students to con gure their scores as scripts, having invested heavily in the
p. 132 ideology of ‘corrective teaching’. (3) The third reason is that there is a sense in which the concept of
distraction comes to Performance Studies from Sociology, where it has a messy, prosaic sense accumulated
from the study of cultural spectacles, or from commercial Research and Development, where it has
immediate worldly and nancial consequences, such as greater military success, safer driving, increased
consumer spending, or their opposites. (4) The fourth reason is the ruling ideology of silent listening and
full attention, which dominates classical music in particular. As this chapter suggests, however, distraction
reminds the performer that silence and attention are not facts but ideologies, not givens but tasks—as is
clear from the music psychology literature on sight-reading, in which ‘distracted inner hearing’ is claimed
to be a reason for mistakes (Wöllner et al. 2003).

Despite these curious absences and the strange reception often accorded to distraction, a shift in C8.P17
contemporary musical culture re ecting global technological change and a re-evaluation of the relationship
between music and citizenship a ords the performer the opportunity to accept the fragmentation of
attention and its ‘becoming-distracted-attention’ as not in itself negative, and to apportion greater value to
music’s dynamic temporality. Attempting to recuperate a positive con guration of distraction is my task
here, and I drift towards the basic phenomenological position into which the performer is put by distraction.
My argument is that the relationship between attention and distraction is asymmetrical: distraction
functions as a temporal switch between attention and distraction, and its impact is less that of an anti-
ergonomic interruption and more that of an involution—a return of the body to itself, and thus a pragmatic
means by which the performer can sustain her engagement with the musical sound that is both her task and
her tool.
Susceptible Subjects C8.S3

Such is the broad context for the phenomenological reduction of distraction in the performer’s listening. I C8.P18
now bracket the notion of distraction as a cultural spectacle and begin to describe the aesthetic function of
distraction. After that I bracket the aesthetic moment of distraction in order to descend to its fundamental
structure, which I analyse in terms of distraction’s functioning as a temporal ‘switch’. This temporal

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
bedrock is Husserlian insofar as it is constituted in temporal terms; but it displaces the classical Husserlian
position by not insisting that noesis and noema have to be joined—indeed, by arguing that the intrusion of
distraction into listening is such that it separates, albeit temporarily, these two parts of the intentional act.

I seek to conduct a phenomenological reduction of performing with speci c reference to the impact of C8.P19
distraction on the performer’s listening; this includes the claim that distraction acts as a temporal ‘switch’.
The reduction involves displacing the classical phenomenological con guration of intentionality, namely
that attention is a ‘ray’ emerging from the pure ego to impart meaning to the object, and revising
subsequent developments such as Heidegger on ‘clearings’ and Merleau-Ponty on ‘mental elds’, all of
p. 133 which employ visual logic to describe attention. This displacement a ords me the opportunity to develop
distraction as a speci cally aural event. That this is necessary is implied by Husserl’s famous statement
about consciousness:

Dazed by the confusion between object and mental content, one forgets that the objects of which C8.P20
we are ‘conscious’, are not simply in consciousness as in a box, so that they can merely be found in
it and snatched at in it; but that they are rst constituted as being what they are for us, and as what
they count as for us, in varying forms of objective intention. (2001: 275)

By this I mean that, for the purposes of the phenomenological reduction, if auditory objects are not ‘in a C8.P21
box’ to be seen and grasped by the performer, then perhaps this is because they are not seen at all, and the
manner in which they are ‘constituted’ needs be approached otherwise; this requires displacing the work
that ‘objective intention’ registers on behalf of cognition. Nancy has made progress in this direction in
Listening. Similar revisions have been proposed elsewhere in the literature on distraction and interruption
(Quinones 2006), notably historicizing distraction and moving beyond the classical (and sometimes
apocalyptic) theories of Benjamin and Adorno, and acknowledging the structure of the contemporary
workplace and the contemporary impact of distraction, which is now, quantitatively and qualitatively, as
much mental as physical (Banbury et al. 2001: 24). As one study notes (Vago and Zeidan 2016: 96), ‘Given
the heavy demand of modern life on cognitive load, managing the onslaught of ongoing sensory and mental
events throughout daily life and improving e ciency of mental processing is of high concern.’

I start the reduction by focusing on the complications that arise from listening’s constitution as C8.P22
simultaneously central to performing (whoever heard of a performer who does not listen?) and a parallel
ongoing process (the performer hears the listener sni ng). My way into the reduction is via the
characteristically seductive noema of sound, as this con gures the relationship between desire and
cognition in listening. To summarize the phenomenological reduction, distraction arises on three levels: (1)
as an aesthetic component of musical experience; (2) as a phenomenological component of sonic
experience; (3) as a perceptual component of auditory experience. This is merely pragmatic, if we accept
that performing has the potential to share pleasure and contribute to the generation of social capital.

In order that distraction be reduced fully and its constitution as a temporal switch be acknowledged, we C8.P23
need to understand the context in which it arises. Distraction is a contextual phenomenon, always related to
an auditory stream into which it intrudes and against which it pushes. Distraction arises and is embodied in
many ways, according to the local context. The detailed typology of distraction mentioned above would
distinguish between multiple modes of emergence (Forster 2013). Indeed, dominated as contemporary life is
by ‘a ects’ (and the distraction that writes a ect onto the body), we might note that, given that ‘[a]ction
occurs from within atmospheres that may be distractedly dwelt within, intentionally shaped, barely noticed
p. 134 or felt as overwhelming presence’ (Anderson 2014: 156), it is pragmatic to con gure temporal events as
coalescing into what Dennett (1991: 101–138) calls ‘multiple drafts’, as a matter of parallel event streams
which are riven with distraction that prevents them from reifying into anything more permanent.
Moreover, not only are there multiple modes of distraction, but not all distraction is distracting. Indeed,
knowledge can be categorized as useful or distracting, according to the task in hand and the nature of the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
distraction (Levitin 2015: 33). In certain situations, ‘interruptions cause arousal and stress levels to elevate
and attention to narrow, resulting in faster performance on simpler [though, it appears, not complex] tasks’
(Ratwani et al. 2006: 372) While, for example, McKenzie con gures the ipside of performance as a black-
and-white matter of failure (‘perform or else fail’), distraction is the grey area in which success (however it
is measured) drifts into failure, and vice versa: ‘perform or else get distracted.’

Before moving on to discuss the impact of distraction on the performer’s listening, I pause to note a C8.P24
potential obstacle facing a phenomenological reduction of distraction. Given that some, if not most,
distraction passes by either almost or entirely unnoticed, and may be non-conscious (since auditory
streaming seems not to require conscious attention) (Macken et al. 2003), with its e ects noted and
rationalized, but the distraction itself long since dispersed and ‘run o ’ into the past (Husserl’s phrase),
how can we recognize and individuate it? One way might be to listen out for distraction in reports of its
embodied residues: in accounts of errors, degradation, disruption, suboptimal performance, changes in
planning (what psychologists term ‘interruption lag’ and ‘resumption lag’) (Trafton and Monk 2007: 114),
self-reports of the performer learning from her mistakes, and so on. It is clear that there many events in
which, although it sounds as if a distraction is present, it is as hard for the listener to verify from outside
whether this is indeed the case as it is for the performer to remember afterwards if what happened had been
a distraction. An example illustrates the range of distraction that works through musical events, and the
nature of the ‘threshold’ between non-distraction and distraction. When, on Friday 9 May 2014 at the Royal
Welsh College of Music and Drama in Cardi , Paul Roberts had nished the spoken part of his keynote
address on ‘The Pianist as Actor’, he sat down to perform the rst of Liszt’s Tre Sonetti del Petrarca S.161
(1839–46). Throughout this poetic performance his body was relatively still and calm, and his hands
remained close to the keyboard—with one momentary exception. Just after the chain of descending
diminished seventh chords in bar 59, he raised his head slightly and rubbed his chin, before continuing on.
Nothing to write home about, possibly. But (1) perhaps this was the embodied trace of distraction that had
entered the performance for him somewhere around this point in the music; (2) perhaps it was nothing
more than a physical itch to which his body was responding in passing; (3) perhaps it was the residue of his
body’s pacing of the movement, re ecting the turning point around the Golden Section in this bar; (4)
perhaps something particular to the sound of the piano in the Weston Gallery on this occasion made his ears
prick up and explore the reverberation of the musical descent; (5) perhaps it was the glimmer of a re ection
on the expressive function of the cadenza at this point in the music. Whatever it was, all of these (and other)
p. 135 possible descriptions refer to distraction that may plausibly have arisen, to what Zbikowski (2011: 179)
calls ‘a cascade of mental images’ that a ects the performer.

The level of distraction and the likelihood of its having an impact on the performer’s listening and thus C8.P25
a ecting her performing can be divided into various factors. These include: the irregularity of a series of
distractions, termed the ‘changing state e ect’ (Banbury et al. 2001: 15); the capacity of the mind to process
information e ciently (Macken et al. 2009); whether the mind is forced to adjust the serial order of how
events are processed (Jones et al. 2010); and the extent to which the distraction violates expectations
(Vachon et al. 2011). Distraction is often more disruptive when it prevents the performer from mentally
‘rehearsing’ the task that has been interrupted (Cades et al. 2007). It often delays the performance of tasks
(Ratwani et al. 2006). It appears to be more prevalent neurologically in the brain’s left hemisphere—the so-
called ‘left-ear advantage’ (Hadlington et al. 2006). It is a function of informational value as well as novelty
(Parmentier et al. 2010). Working memory span is an important factor in how individuals respond to
distraction in the middle of a task and how they deal with the excess information presented by the
distraction (Beaman 2004: 1107). Cognitive ‘load’ in uences the accuracy of task performance (Forster
2013). Distraction can arise from the fact that cognition is required to deal with ‘competing streams’ of
sensation (Ponjavic-Conte et al. 2013). At the ‘threshold’ of conscious listening there is a ‘stochastic
nonlinear bifurcation in neural activity’ (Sergent and Dehaene 2004), which seems to increase the
likelihood of distraction. Banbury et al. (2001: 26-27) conclude that there are eight ways in which sounds

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
can disrupt cognitive processing: (1) acoustic change, mediated by the sound’s perceptual organization
(repeated sounds are not disruptive); (2) change in pitch/timbre/tempo but not sound level; (3) non-speech
sounds; (4) irrelevant sounds, not necessarily simultaneously with what they interrupt; (5) habituation to
interruptions is rare; (6) impact on short-term memory, particularly seriation; (7) high-level tasks can be
disrupted by semantic properties of the sound; and (8) disruption caused by con ict of concurrent cognitive
processes.

Given all of the above, we should not be surprised that cognition is riven with distraction. In fact, distraction C8.P26
should be expected, and Husserl’s focus on the ‘accomplishment’ of intentional acts needs to be displaced
by focusing on the types of event that he found di cult to describe, or which he simply believed were
marginal to consciousness: interruptions, delays, incomplete intentions, mistakes—and distractions.
Ganeri (2017), for example, is clear about the priority, reversing the standard con guration between
‘attention’ and ‘self’. Attention, he argues, with its two roles of ‘placing-on’ and ‘focusing-at’, precedes the
self and has greater explanatory power. Marder (2009: 31) takes a similar position, analysing attention as a
‘proto-intentionality’. Indeed, the priority of attention (though I prefer to stretch Ganeri’s term to embrace
distraction or to talk of ‘distracted attention’) over self is obvious from the millions of little musical events
in which distraction does not undo or destroy self but complicates and distributes it. To pause over one
example: in the Queen Elizabeth Hall in London on Tuesday 7 October 2003, during a performance of
Xenakis’s Eonta (1963–64) for piano and ve brass instruments (the rst work on the programme), the
p. 136 pianist Nicolas Hodges pushed his glasses back up his nose during the extended opening solo section of
the work several times. He seemed to nd the time to do so despite the extreme technical demands of the
music, which, constructed stochastically, is hardly idiomatic for the piano. It seemed that were he not to
have nudged his glasses each time, then what was probably a slight distraction on the boundary between
non-conscious and conscious thought might have become a fully conscious and threatening distraction,
and a real musical problem for the continuity of the music, which in the rst section has no rests or
equivalent places for either hand to move away from the keyboard.

In this context, how does listening get con gured? Listening regimes—part social, part cognitive— C8.P27
con gure listening as either productive-ergonomic or uid-associative. The two are not mutually
exclusive: structural listening is located towards one end of the spectrum, the music appreciation racket lies
in the middle, and towards the other end of the spectrum lies what Dahlhaus calls ‘trivialised listening’.
Many listening regimes attempt to minimize distraction, privileging attention on the phenomenological
assumption that it functions as the ‘glue’ holding together subject and object (Marder 2011); at its most
extreme, this assumption is the bourgeois fear of the dispersion of subjectivity into the sonic environment,
lost but not found, the fear that our bodies are doomed to being inhabited by a ‘relentless monkey mind’
(Vago and Zeidan 2016: 96). For example, in the space between music appreciation and structural listening,
it has often been assumed that the most productive listening concentrates on the relation of each moment
to the larger whole and ignores distraction. Roberts assumes this in his statement about ‘blocks of wood’
cited above. Looser accounts of listening that accept the inevitability of distraction and their
phenomenological constitution as shock, surprise, and other such phenomena (Love 2008; Nancy 2000)
assume various forms: from strong musical experiences in which, as one listener confesses, ‘I felt a bliss
without bounds and “ oated”, so it seemed me, several meters above the oor. I had a vision of another,
better life—a feeling of there being unimagined possibilities even for me’ (Gabrielsson 2011: 156) to the
reports of listeners ‘who may or may not have been, at the time, thinking of something else entirely’
(Godlovitch 1998: 127), from listeners who use music as a means of depersonalizing and de-realizing their
everyday experience (Herbert 2011a) to music ‘self-chosen’ for its ability to a ord ‘distraction, energizing,
entrainment, and meaning enhancement’ (Lamont et al. 2016)—whether washing up, wishing, or working.

One important point to note about distraction’s impact on the performer’s listening is that, contrary to the C8.P28
fear of many theories, distraction does not weaken or destroy the ‘adequation’ of the performer and the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
musical object. In fact, it reinforces this reciprocity by engaging the performer in an act of ‘perceptual
learning’ that draws the loss of self around full circle towards the found self (Clarke 2014). The separation of
noesis and noema already mentioned is temporary (to deploy the term ‘switch’ as a metaphor, it is as if the
performer’s intentional act is switched o and on again). In this respect, distraction can be described with
the terms of ecological psychology, emphasizing the environmental presence of sound (internal processes
account only for part of what happens). Consider Clarke’s phrasing in Ways of Listening (2005: 31):

p. 137 The environmental events that gave rise to the particular connections and weightings in the C8.P29
system (or the synaptic links in the brain) are manifest relationships in the concrete physical
world. The subsequent ‘tuning’ of the network (whether arti cial model or actual brain) is the
result of exposure to those real events—their trace, or residue. That trace, and its reactivation, is
experienced as a dynamic state of the network and thus a state of mind—an awareness of real-
world relationships.

This is a neat way of making the point that, while distraction is tethered to sound in the real world C8.P30
(‘manifest relationships’), it loosens up listening (‘retuning’ and re-adaptation of the system) and
challenges the performer to make something of what is happening right now (assimilating the ‘trace or
residue’ into the event).

Temporal Switches C8.S4

In relation to the matter of music performance, distraction might seem to be an anomalous subject. Surely it C8.P31
gets in the way of e cacious, e cient, and e ective performing (McKenzie’s three ways of evaluating
contemporary events). Surely it is only a peripheral, minor phenomenon that disappears upon better
listening. There is some truth in these two claims. For distraction is cunning: it does cause listening to
switch direction, it does cause listening to re ect upon itself, it does not create time as such, and it does not
switch listening to a known or predictable second time. In this respect, distraction seems to ‘delay’
performing: it expands time, multiplies it, disperses it (hence distraction is deplored in many of the music
discourses that invest heavily in attentive listening), and opens up musical time to indeterminacy.
Nevertheless, it is acknowledged that ‘[m]usic communicates critical time dimensions into our perceptual
processes’ (Thaut 2008: 34), and one way in which this happens is through distraction. Here, nally, after
the aesthetic register of distraction discussed in the previous section, the phenomenological reduction
reaches its core argument: distraction is a temporal switch. How best to describe this mechanism?

Distraction is a temporal event. It is rooted in a temporality concerned with the experience of a never fully C8.P32
present time, distensio. For St Augustine, this describes the distracted soul. In the contemporary world, we
might note that ‘consciousness can be modulated by focal attention’ (Edelman 2005: 61) and that it
mobilizes a shared intentionality. This mobilizing is a beginning and a switching. Indeed, distraction is
always a beginning, an injection of energy into the system. Distraction springs up out of events, and,
notwithstanding its psychological inscription onto ‘retention’ (and later assimilation by the relevant
listening regime), its focus is always on what comes next, on the future, which is ‘the fundamental
phenomenon of time’ (Heidegger 1992: 14). In this sense, distraction qua temporal switch does not arise
within individual musical events, but between events, ‘constituting’ these very changes. Husserl (1983: 222)
p. 138 speaks of ‘attentional changes’, but distraction is only partly a matter of a shifting perspective on the
same object. It also concerns the nature of an object being opened up not just for validation of a mental
schema but, more radically, for the immediate future development of listening. Although normally ‘the
purely sensory character of perception is usually not at all evident to a perceiver’ (Clarke 2005: 32),
distraction makes the sensory character of listening explicit, switching the performer back towards herself,
her body being ‘folded back onto itself’ (Nancy 2007: 17, 18, 30, 37, 41). In this manner, she is no longer

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
transparent to the music and its meaning, no longer an empty vessel for its emergence. Performer and
object are set in motion around each other and generate ‘resonance’, the temporal switch plays itself out as
they temporarily resist each other’s embrace, and the performer is forced to recall the basic principle that
‘the primary function of auditory perception is to discover what sounds are the sound of, and what to do
about them’ (Clarke 2005: 3).

Distraction acts like a trap set for listening by temporality, a trap for the ‘natural attitude’ (Husserl). It is a C8.P33
temporal device that switches the very strength of attention (its determined force for drawing together the
performer and the work) into its weakness. This switch in energetic investment opens up a temporal
discontinuity, and writes a new ‘draft’ of consciousness that recon gures the virtual body of listening by
returning the listener’s subjectivity and the auditory object to their rightful domains alongside, but not
overlapping with, each other. The switch loosens up the ‘intentional overreaching’ that follows the
phenomenological ‘pairing’ of listener and sound (Husserl 1973: 112); thus does distraction retain a certain
contingency and indeterminacy in the event at the cost of a quicker accomplishment of transcendental
subjectivity. In this sense, distraction reminds the performer that attention is a partial, albeit seductive,
model for her listening, which can easily trick her into thinking that attention is natural and that the mind is
the only thing contributing to the listening process.

So distraction has the e ect of both separating attention from its object and drawing the two closer to each C8.P34
other. Such is its cunning and the seductiveness of the sonic noema that the performer’s consciousness is
dismantled and reconstructed, and the performer is reminded of music’s ‘capacity to e ect shifts of
consciousness that support an individual’s sense of daily psychological balance’ (Herbert 2011b: 306).
Following McKenzie’s judicious conclusion (2001: 177), distraction reveals that ‘the truth of everything
emerges through the cracked and uneven joints that bind together performatives and performances’. These
‘cracks’ and ‘unevenness’ reveal a Nietzschean cunning to the way in which distraction plants switches
within listening and cajoles the listener into confessing, ‘We are a plurality that has imagined itself a unity’
(Lingis 1985: 39) and coming clean about the ideological basis of certain listening regimes.

The beginning that distraction opens up in listening is neatly captured by Marder’s critique of the standard C8.P35
opposition between attention and distraction: ‘Not only does the act of attending animate or enliven
consciousness in the passage from inactional and indeterminate potentiality to the actional determination
of a noema but it also coincides with intentionality, itself the form of life proper to consciousness’ (Marder
2009: 29). This is congruent with the function of distraction as described by the phenomenological
p. 139 reduction: distraction and attention, set in motion by the distraction’s temporal switch, work alongside
each other to ‘animate’ listening and set it in motion towards an appropriate listening regime. What I term
‘distraction’ is what Marder (p. 33) terms an ‘attentive animation of consciousness’ or ‘the attentional
regard’:

Attention requires some measure of distraction, but this is not to say that it is indistinguishable C8.P36
from its opposite. The attentional animation of consciousness must be recognized for what it is:
the rst, albeit relentlessly repeated, demarcation of the intentional eld or a pre-interpretation
of the noema that demands a further deepening and, perhaps, a di erent kind of attention (call it
‘attentive,’ as opposed to ‘attentional’) capable of tarrying along with the intended singularity.
Although the enlivening of consciousness by the attentional regard is a necessary precondition for
egoic life, this life cannot be reduced to its initial moment, which opens up a passage to the
actional mode but does not entirely coincide with it.

This is an essential point of the phenomenological reduction: distraction is the ‘ rst demarcation’ of C8.P37
consciousness, a switch that opens up the temporality of listening. And thus ‘To pay attention’, as Russon
(2017: 30) adapts Merleau-Ponty, ‘is to make oneself answerable to the demands of the object’—precisely,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
to give oneself up to distraction. Moreover, given that ‘the brain is never truly resting’ (Vago and Zeidan
2016: 97), attention and distraction must be con gured as interweaved. If the brain is never in a zero-
degree state, then they cannot be opposing forces, attention being some kind of positive force owing away
from a resting state and distraction being a negative force owing in the opposite direction, since neither
attention nor distraction ‘starts’ switching from a zero degree cost-neutral base.

Distraction’s temporal switch makes it di cult to disentangle which moments of listening are grounded in C8.P38
attention and which in distraction: ‘music has the capacity to call into question the rigid separation of
subject and object and to play with states which do not wholly conform with either the one or the other’
(Clarke, 2005: 86–87). Good examples of this di culty of disentangling distraction and attention—it being
more productive to speak of ‘distracted attention’—occurred during Nelson Goerner’s recital in the
Bridgewater Hall in Manchester on Friday 18 March 2005. His programme consisted of the following works:
Bach-Brahms’ Chaconne from Partita No. 2 in D minor BWV 1004, Beethoven’s Sonata No. 31 in A♭ major Op.
110 (1821), Granados’ Goyescas No. 2 (1909–11), and Debussy’s Estampes L.108 (1903). His three encores were
Chopin’s Nocturne in C minor Op. 48 No. 1 (1841), Chopin’s nal Prelude No. 24 in D minor Op. 28 (1935–
39), and Liszt’s Grandes Études de Paganini No. 2 S. 141 (1851). Goerner hardly looked up from the keyboard at
all during the recital, spending almost all the time poring over his hands or staring straight ahead. So the
few heavenward glances may be signi cant: in the Beethoven there was one at the start of the G minor
arioso when it recapitulates the A♭ minor statement; in the Granados there were two, the rst at the general
pause, the second in the coda when there is a noticeable shift of harmony; in the second Estampe there were
two, the rst during the distant empty texture right at the start, the second at the end when this texture
p. 140 returns; in the third Estampe there was a look up at the ceiling when the melody rst enters above the
tremolo accompaniment. These head movements (1) may re ect the residues of what Cha n et al. (2002)
term ‘performance cues’; (2) may have some kind of personal signi cance for Goerner’s interpretations of
the works; (3) may be distant echoes of his structural analysis (however loosely or intuitively conducted) of
the works; (4) may have arisen because his neck was aching; (5) may have been responses to coughing in the
audience just before these moments; (6) may be part of Goerner’s performing style, a harmless mannerism
used in public in certain pieces (there were no such glances during the encores); (7) may be a combination of
the above. Goerner’s ve head movements may also have been the embodied response in the only free part
of his body to distractions perceived either within the hall or within his own body and its mind, and may
themselves have been distractions forcing Goerner’s attention to switch from one thing to another. They
were certainly switches in the performance event that segmented and shaped his performing.

In this respect, supervenient upon the phenomenology of distraction is a rhetoric of switches. This rhetoric C8.P39
can be deployed on behalf of various claims: to give two historical examples, consider the Platonic attempt
to ban music from the city for fear of its power over the polis, or the Enlightenment fear of noise and the
historical silencing of the concert hall, both of which subscribe to the same logic. However, distraction is
cunning and resists attention’s claim to have the nal word on listening and on what meanings performing
can a ord. In contrast to theories of listening that fetishize attention, distraction allows the performer to
judge events only by their auditory e ects, by how sounds sound, without fetishizing the search for a
common measure between sounds—‘Just an attention to the activity of sounds.’ (Cage 1961: 10) The real
moral of distraction is this: distraction is as strong a force for listening as attention, and, in fact, the two are
co-constituting (the phenomenological reduction concludes that ‘distraction’ should be displaced by the
term ‘distracted attention’). Thus, while distraction cannot function as the ground for a performance
practice or an aesthetic (because, in Husserlian terms, it cannot be ‘intuited’, i.e. perceived fully)—that is to
say, while it cannot be the telos of listening activity and performing decisions (even Cage’s aesthetic is not
grounded on distraction, but simply acknowledges it, albeit more willingly than most)—the rhetoric of
distraction nevertheless allows the performer to become more aware of distraction and more cognisant of
its potential happening.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
Loosening Up C8.S5

We might ask why the discourses supporting performing music still focus on (at best) ignoring the temporal C8.P40
switches that it presents and (at worst) eradicating distraction, given that doing so only constrains the
performer to what she had imagined would happen on stage. The question is why they insist on training the
p. 141 performer not to relinquish too much agency like Cage, not to pay attention to external noises like Hamelin,
not to wear loose- tting glasses like Hodges, not to think too much (or in a particular way) like Roberts,
not to move in what might be construed as an extraneous manner like Goerner—as if simply not doing these
things and instead remaining bound to a legalistic con guration of the concept of attention is a guarantee of
success. What might it mean to con gure performing as a task that includes the realization of distraction’s
potential to be a creative moment of the performer’s listening and central to the generation of pleasure and
the accumulation of social capital? What consolation is there for the performer?

Certainly, drifting from one extreme (the standard position, which is still prevalent pedagogically even C8.P41
though research has moved on)—in which performing is constrained to a logic of representation and the
performer’s cognition is measured in terms of its validation of centrally processed and validated schemas—
does not automatically imply a direction of travel towards the other extreme, in which, as one playwright
claims (Foreman 1978: 131) in Cageian fashion, ‘The text = strategies for allowing the world to interfere.’
‘Mindfulness’ and ‘mind wandering’ (the latter is distraction’s cousin) are not opposites and not mutually
exclusive (Vago and Zeidan: 2016). Distraction does not impact upon only one type of listening. Its temporal
switches work through listening in all its registers, including the representational as much as the
imaginative, the sensuous-embodied as much as the critical-evaluative. The switches are always tethered to
sound, to sense data, to what the performer (thinks that she) hears, which is why the ‘wavering listening’
proposed by Szendy (2008: 103, 104, 122) is useful but constrained by its subscription to listening-by-
association. And distraction’s switches tether the performer to the sounds she is listening to, reminding her
that, notwithstanding the virtual gestures created by the sounds qua tones, she remains physically within
the world, which is why the listening espoused in Beyond Structural Listening (Dell’Antonio 2004) is not
forceful enough, for, notwithstanding its rethinking of ‘structure’, it still subscribes to a discourse of
expertise.

More broadly, if McKenzie’s prophetic claim (2001: 176, bold in original) that ‘Performance will be to the C8.P42
twentieth and twenty- rst centuries / what discipline was to the eighteenth and nineteenth: / an onto-
historical formation of power and knowledge’ is to have teeth, then it needs to give distraction a central
position, a ord distraction and its switches the opportunity to work through listening, and con gure this
working through as a signi cantly more vital phenomenon than just the ipside qua failure of attention.
Indeed, North’s whole project in The Problem of Distraction is to argue against con guring distraction and
attention as merely opposites. He argues (2011: 12, 13) that distraction is more than ‘an anthropological
matter’: it is ‘a paraontological relationship of thought to non-being and its variants: not-quite-being,
more-than-being, not-yet-being, no-longer-being.’ So, pace McKenzie, distraction is what draws
‘discipline’ into ‘performance’. Alongside a logic of ‘perform—or else’, we need a logic of distracted
listening. And we need to incorporate distraction more centrally into our engagement with noise, sound,
and music (including how switches alert us to changes from e.g. noise to sound, or vice versa). As the
planet’s population continues to grow uncontrollably and to concentrate itself in a smaller number of
p. 142 inhabitable places, how to survive distraction will become a key human problem, even more so than is
already panic-theorized by writers like Pettman (2015).

In fact, the task will be to learn, less how to survive distraction, and more how to live with it: ‘I no sooner C8.P43
start to work than the telephone rings’ (Cage 1967: 99) In this respect, more aural training will only help the
performer, for it develops key transferable skills in her auditory domain: it develops a familiarity with

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
temporal switches and a con dence in apprehending their varied modes of intrusion into her listening. As
Vago and Zeidan (2016: 107–108) conclude, ‘A sense of peace and quiet in the mind is proposed to arise
through mental training in concentration, nonconceptuality, and discernment, in contrast to the untrained
frenetic restlessness of mental time travel that is characteristic of daily activity in the postmodern setting.’
By developing aural skills (conceived broadly beyond the imitation and reproduction tasks that quite
understandably dominate educational curricula), the performer’s ability to react prudently to distraction
develops in tandem. Cage (1979: 180–181), once again, seems to have been an early adopter of this idea,
re ecting on ‘The Future of Music’ as follows:

But if there is any experience more than another which conduces to open-mindedness, it is the C8.P44
experience of being bothered by another, of being interrupted by another. ‘We are studying being
interrupted’. Say we do not practice spiritual discipline. The telephone then does it for us. It opens
us to the world ‘outside’.

This is, perhaps, the moral of Ishiguro’s The Unconsoled. As he gets closer to the moment of stepping on C8.P45
stage to perform, Ryder gradually comes to acknowledge that he may never be fully in control either
musically or socially, and that this does not actually matter: performing is not solely about control, even in
conservatoires. The multiple streams or ‘drafts’ of listening cannot be reduced to a single input–output
function managed by a central director, and distraction will always potentially arise as and when it will,
unseating the best planned scenes and loosening up prepared gestures. Ryder re ects on his performing
thus (Ishiguro 1995: 356–357):

After a short moment to collect my thoughts I went into the vertiginous opening of Asbestos and
Fibre. Then as the rst movement settled into its more re ective phase, I became increasingly
relaxed, so much so that I found miself playing most of the rst movement with my eyes closed.

As I began the second movement, I opened my eyes again and found the afternoon sunshine
streaming through the window behind me, throwing my shadow sharply across the keyboard. Even
the demands of the second movement, however, did nothing to alter my calm. Indeed, I realized I
was in absolute control of every dimension of the composition. I recalled how worried I had
allowed myself to become over the course of the day and now felt utterly foolish for having done
so. Moreover, now I was in the midst of the piece, it seemed inconceivable that my mother would
not be moved by it. The simple fact was, I had no reason whatsoever to feel anything other than
p. 143 utter con dence concerning the evening’s performance.

It was as I was entering the sublime melancholy of the third movement that I became aware of a
noise in the background. At rst I thought it was connected with the soft pedal, and then that it was
something to do with the oor. It was a faint, rhythmic noise that would stop and start, and for
some time I tried not to pay any attention to it. But it continued to return, and then, during the
pianissimo passages mid-way through the movement, I realized that someone was digging outside
not far away.

As he performs, Ryder realizes an essential truth about performing and the impact of distractions upon his C8.P47
listening: ‘What is being fooled? Not the ear but the mind’ (Cage 1967: 109).
So what should the performer do, given the omnipresent potential of distraction intervening in her C8.P48
listening? She should breathe, trust her ears, and engage all her senses equally. For her performing is open
to distraction precisely because in the contemporary world expertise and intentionality have been
recon gured (downgraded, perhaps, in the body’s list of priorities), and listening has come to take the top
spot. Darwin redux: survival of the best listeners. This is not to say that anything goes for the performer as
she listens. It is simply that something comes to pass. Performing in general should now be entirely future-
facing, open to what may happen, and no longer concerned either with returning to close the backwards-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
facing circle of representative delity or with fetishizing technical adequation to a set of mental projections
—with assuming that, even if one or both of these intentions are necessary (according to the musical style
of the piece being performed), that this automatically also makes the intention su cient.

What might be lost, then, from not acknowledging the impact of distraction on the performer’s listening? C8.P49
From not acknowledging that attention is in fact always distracted attention? From not acknowledging that
the temporal switches of attention driven by distraction could be productive both within and beyond
performing? There will, after all, always be someone digging about noisily nearby, whether in a ctional
European city or a stone’s throw from the Wigmore Hall, and both Ryder and Hamelin have nothing to lose
from acknowledging that their performing takes place in the world. Aural training for the performer is at the
heart of her development, and, done carefully, it can teach her, not to listen less or more, but to listen more
patiently and openly to what happens while she is performing. She may be delayed, but if the performer
opens her ears then she may just nd herself marvelling at what happens.
References C8.S6

Adorno, T. (1991). On the fetish character in music and the regression of listening. In The culture industry: Selected essays on mass C8.P50
culture, 29–60. Routledge.
Google Scholar Google Preview WorldCat COPAC

Anderson, B. (2014). Encountering a ect: Capacities, apparatuses, conditions. Ashgate. C8.P51

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Banbury, B., Macken, W., Tremblay, S., and Jones, D. (2001). Auditory distraction and short-term memory: Phenomena and C8.P52
practical implications. Human Factors 43(1): 12–29.
Google Scholar WorldCat

p. 144 Beaman, P. (2004). The irrelevant sound phenomenon revisited: What role for working memory capacity? Journal of C8.P53
Experimental Psychology: Learning, Memory, and Cognition 30(5): 1106–1118.
Google Scholar WorldCat

Beaman, P., and Holt, N. (2007). Reverberant auditory environments: The e ects of multiple echoes on distraction by ʻirrelevantʼ C8.P54
speech. Applied Cognitive Psychology 21: 1077–1090.
Google Scholar WorldCat

Benjamin, W. (2008). The work of art in the age of its technological reproducibility. In The work of art in the age of its C8.P55
technological reproducibility and other writings on media, 19–55. Harvard University Press.
Google Scholar Google Preview WorldCat COPAC

Cades, D., Tra on, J. G., Boehm Davis, D., and Monk, C. (2007). Does the di iculty of an interruption a ect our ability to resume? C8.P56
Proceedings of the Human Factors and Ergonomics Society 51st Annual Meeting 51(4): 234–238.
Google Scholar WorldCat

Cage, J. (1961). Experimental music. In Silence: Lectures and writings, 7–12. Wesleyan University Press. C8.P57
Google Scholar Google Preview WorldCat COPAC

Cage, J. (1967). Juilliard Lecture. In A year from Monday, 95–111. Wesleyan University Press. C8.P58
Google Scholar Google Preview WorldCat COPAC

Cage, J. (1979). The future of music. In Empty words: Writings ʼ73–ʼ78, 177–187. Wesleyan University Press. C8.P59
Google Scholar Google Preview WorldCat COPAC

Cha in, C., Imreh, G., and Crawford, M. (2002). Practicing perfection: Memory and piano performance. Erlbaum. C8.P60
Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (2005). Ways of listening: An ecological approach to the perception of musical meaning. Oxford University Press. C8.P61
Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (2014). Lost and found in music: Music, consciousness and subjectivity. Musicae Scientiae 18(3): 354–368. C8.P62
Google Scholar WorldCat

Clayton, M., Sager, R., and Will, U. (2005). In time with the music: the concept of entrainment and its significance for C8.P63
ethnomusicology. European Meetings in Ethnomusicology 11: 3–142.
Google Scholar WorldCat

Crawford, K. (2009). Following you: Disciplines of listening in social media. Continuum 23(4): 525–535. C8.P64
Google Scholar WorldCat
DellʼAntonio, A. (ed.) (2004). Beyond structural listening? Postmodern modes of hearing. University of California Press. C8.P65
Google Scholar Google Preview WorldCat COPAC

Dennett, D. (1991). Consciousness explained. Little, Brown. C8.P66


Google Scholar Google Preview WorldCat COPAC

Edelman, G. (2005). Wider than the sky: A revolutionary view of consciousness. Penguin Books. C8.P67
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
Foreman, R. (1978). 14 things I tell myself when I fall into the trap of making the writing imitate ʻexperienceʼ. Semiotext(e) 3(2): C8.P68
124–132.
Google Scholar WorldCat

Forster S. (2013). Distraction and mind-wandering under load. Frontiers in Psychology 4: 283. C8.P69
https://doi:10.3389/fpsyg.2013.00283
Google Scholar WorldCat

Gabrielsson, A. (2011). Strong experiences with music: Music is much more than just music. Oxford University Press. C8.P70
Google Scholar Google Preview WorldCat COPAC

Ganeri, J. (2017). Attention, not self. Oxford University Press. C8.P71


Google Scholar Google Preview WorldCat COPAC

Godlovitch, S. (1998). Musical performance: A philosophical study. Routledge. C8.P72


Google Scholar Google Preview WorldCat COPAC

Hadlington, L., Bridges, A., and Beaman, P. (2006). A le -ear disadvantage for the presentation of irrelevant sound: Manipulations C8.P73
of task requirements and changing state. Brain and Cognition 61: 159–171.
Google Scholar WorldCat

Heidegger, M. (1992). The concept of time (trans. W. McNeill). Blackwell. C8.P74


Google Scholar Google Preview WorldCat COPAC

Herbert, R. (2011a). Everyday music listening: Absorption, dissociation and trancing. Ashgate. C8.P75
Google Scholar Google Preview WorldCat COPAC

Herbert, R. (2011b). Consciousness and everyday music listening: Trancing, dissociation and absorption. In D. Clarke and C8.P76
E. Clarke (eds), Music and consciousness: Philosophical, psychological, and cultural perspectives, 295–308. Oxford University
Press.
Google Scholar Google Preview WorldCat COPAC

Home-Cook, G. (2015). Theatre and aural attention: Stretching ourselves. Palgrave Macmillan. C8.P77
Google Scholar Google Preview WorldCat COPAC

p. 145 Husserl, E. (1973). Cartesian meditations (trans. D. Cairns). Martinus Nijho . C8.P78
Google Scholar Google Preview WorldCat COPAC

Husserl, E. (1983). Ideas pertaining to a pure phenomenology and to a phenomenological philosophy, vol. 1 (trans. F. Kersten). C8.P79
Martinus Nijho .
Google Scholar Google Preview WorldCat COPAC

Husserl, E. (2001). Logical investigations (trans. J. Findlay). Routledge. C8.P80


Google Scholar Google Preview WorldCat COPAC

Huxley, A. (1958). Brave new world revisited. Harper & Row. C8.P81
Google Scholar Google Preview WorldCat COPAC

Ishiguro, K. (1995). The unconsoled. Faber & Faber. C8.P82


Google Scholar Google Preview WorldCat COPAC

Jones, D., Hughes, R., and Macken, W. (2010). Auditory distraction and serial memory: The avoidable and the ineluctable. Noise C8.P83
Health 12(49): 201–209.
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
Kracauer, S. (1995). Cult of distraction: On Berlinʼs picture palaces. In S. Kracauer, The Mass Ornament: Weimar Essays (trans. C8.P84
T. Levin), 323–328. Harvard University Press.
Google Scholar Google Preview WorldCat COPAC

Lamont, A., Greasley, A., and Sloboda, J. (2016). Choosing to hear music: Motivation, process, and e ect. In S. Hallam, I. Cross, C8.P85
and M. Thaut (eds), The Oxford handbook of music psychology (2nd edn), 711–724. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Lavie, N., Hirst, A., and de Fockert, J. (2004). Load theory of selective attention and cognitive control. Journal of Experimental C8.P86
Psychology: General 133(3): 339–354.
Google Scholar WorldCat

Levitin, D. (2015). The organized mind: Thinking straight in the age of information overload. Penguin. C8.P87
Google Scholar Google Preview WorldCat COPAC

Liebl, A., Haller, J., Jödicke, B., Baumgartner, H., Schlittmeier, S., and Hellbrück, J. (2012). Combined e ects of acoustic and C8.P88
visual distraction on cognitive performance and well-being. Applied Ergonomics 43: 424–434.
Google Scholar WorldCat

Lingis, A. (1985). The will to power. In D. Allison (ed.), The New Nietzsche, 37–63. MIT Press. C8.P89
Google Scholar Google Preview WorldCat COPAC

Love, K. (2008). Being startled: Phenomenology at the edge of meaning. PhaenEx 3(2): 149–178. C8.P90
Google Scholar WorldCat

Lyotard, J.-F. (1989). Judiciousness in dispute, or Kant a er Marx. In A. Benjamin (ed.), The Lyotard Reader (trans. C. Lindsay), C8.P91
324–359. Blackwell.
Google Scholar Google Preview WorldCat COPAC

Macken, W., Phelps, F., and Jones, D. (2009). What causes auditory distraction? Psychonomic Bulletin & Review 16(1): 139–144. C8.P92
Google Scholar WorldCat

Macken, W., Tremblay, S., Houghton, R., Nicholls, A., and Jones, D. (2003). Does auditory streaming require attention? Evidence C8.P93
from attentional selectivity in short-term memory. Journal of Experimental Psychology: Human Perception and Performance
29(1): 43–51.
Google Scholar WorldCat

Marder, M. (2009). What is living and what is dead in attention? Research in Phenomenology 39: 29–51. C8.P94
Google Scholar WorldCat

Marder, M. (2011). Phenomenology of distraction, or attention in the fissuring of time and space. Research in Phenomenology 41: C8.P95
396–419.
Google Scholar WorldCat

McKenzie, J. (2001). Perform or else: From discipline to performance. Routledge. C8.P96


Google Scholar Google Preview WorldCat COPAC
Montaigne, M. de (2003). Essais (trans. D. Frame). Everymanʼs Library. C8.P97
Google Scholar Google Preview WorldCat COPAC

Nancy, J.-L. (2000). The surprise of the event. In Being singular plural (trans. R. Richardson and A. OʼByrne), 159–176. Stanford C8.P98
University Press.
Google Scholar Google Preview WorldCat COPAC

Nancy, J.-L. (2007). Listening (trans. C. Mandell). Fordham University Press. C8.P99

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Nettl, B. (2005). The study of ethnomusicology: Thirty-one issues and concepts (2nd edn). University of Illinois Press. C8.P100
Google Scholar Google Preview WorldCat COPAC

North, P. (2011). The problem of distraction. Stanford University Press. C8.P101


Google Scholar Google Preview WorldCat COPAC

OʼCallaghan, C. (2007). Sounds: A philosophical theory. Oxford University Press. C8.P102


Google Scholar Google Preview WorldCat COPAC

Parmentier, F., Elsley, J., and Ljungberg, J. (2010). Behavioral distraction by auditory novelty is not only about novelty: The role C8.P103
of the distracterʼs informational value. Cognition 115(3): 504–511.
Google Scholar WorldCat

Pettman, D. (2015). Infinite distraction: Paying attention to social media. Polity. C8.P104
Google Scholar Google Preview WorldCat COPAC

p. 146 Phillips, N. (2016). Distraction: Problems of attention in eighteenth-century literature. Johns Hopkins University Press. C8.P105
Google Scholar Google Preview WorldCat COPAC

Ponjavic-Conte, K., Hambrook, D., Pavlovic, S., and Tata, M. (2013). Dynamics of distraction: Competition among auditory C8.P106
streams modulates gain and disrupts inter-trial phase coherence in the human electroencephalogram. PloS ONE 8(1): e53953.
https://doi.org/10.1371/journal.pone.0053953
Google Scholar WorldCat

Quinones, M. (2006). Towards a musicology of distraction. Conference presentation, SIBE 2006, ʻExperiència musical, cultura C8.P107
globalʼ, March, Palma de Mallorca.
Google Scholar Google Preview WorldCat COPAC

Ratwani, R., Tra on, J. G., and Myers, C. (2006). Helpful or harmful? Examining the e ects of interruptions on task performance. C8.P108
Proceedings of the Human Factors and Ergonomics Society 50th Annual Meeting, 372–375.
https://doi.org/DOI:10.1177/154193120605000334
Google Scholar Google Preview WorldCat COPAC

Roberts, P. (1996). Images: The piano music of Claude Debussy. Amadeus Press. C8.P109
Google Scholar Google Preview WorldCat COPAC

La Rochefoucauld, F. de (1959). Maxims (trans. L. Tancock). Penguin. C8.P110


Google Scholar Google Preview WorldCat COPAC

Russon, J. (2017). Freedom and passivity: Attention, work, and language. In K. Jacobson and J. Russon (eds), Perception and its C8.P111
development in Merleau-Pontyʼs phenomenology, 25–39. University of Toronto Press.
Google Scholar Google Preview WorldCat COPAC

Sergent, C., and Dehaene, S. (2004). Is consciousness a gradual phenomenon? Evidence for an all-or-none bifurcation during the C8.P112
attentional blink. Psychological Sciences 15(11): 720–728.
Google Scholar WorldCat

Szendy, P. (2008). Listen: A history of our ears (trans. C. Mandell). Fordham University Press. C8.P113
Google Scholar Google Preview WorldCat COPAC

Thaut, M. (2008). Rhythm, music, and the brain: Scientific foundations and clinical applications. Routledge. C8.P114
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469418 by National Science & Technology Library user on 26 May 2023
Tra on, J. G., and Monk, C. (2007). Task interruptions. Reviews of Human Factors and Ergonomics 3(1): 111–126. C8.P115
Google Scholar WorldCat

Vachon, F., Hughes, R., and Jones, D. (2011). Broken expectations: Violation of expectancies, not novelty, captures auditory C8.P116
attention. Journal of Experimental Psychology: Learning Memory and Cognition 38(1): 164–177.
Google Scholar WorldCat

Vago, D., and Zeidan, F. (2016). The brain on silent: Mind wandering, mindful awareness, and states of mental tranquillity. Annals C8.P117
of the New York Academy of Sciences 1373: 96–113.
Google Scholar WorldCat

Wöllner, C., Halfpenny, E., Ho, S., and Kurosawa, K. (2003). The e ects of distracted inner hearing on sight-reading. Psychology of C8.P118
Music 31(4): 377–389.
Google Scholar WorldCat

Wood, M. (2011). The habits of distraction. Sussex Academic Press. C8.P119


Google Scholar Google Preview WorldCat COPAC

Zbikowski, L. (2011). Music, language, and kinds of consciousness. In D. Clarke and E. Clarke (eds), Music and consciousness: C8.P120
Philosophical, psychological, and cultural perspectives, 179–192. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
CHAPTER

9 Music, Evolution, and the Experience of Time  C9

John C. Bispham

https://doi.org/10.1093/oxfordhb/9780190947279.013.14 Pages 149–C9.P154


Published: 08 December 2021

Abstract
This chapter explores the psychological structure and perception of time in music in light of recent
theoretical discussion and proposals for speci c features of the human faculty for music—qualities
that are at once universally present and operational in music across cultures whilst also being unique
to our species and to the domain of music. The author contends that music’s architectural foundations
—con gurations of musical pulse, musical tone, and musical motivation—provide a sustained
attentional structure for managing personal experience and interpersonal interaction and o er a
continually renewing phenomenological link between the immediate past, the perceptual present and
future expectation. A crucial and distinguishing feature of our experience with music is thus the
particular way in which we share intersubjective time and enact and create an extended moment by
constantly merging from one perceptual present into the next.

Keywords: music, time, evolution, pulse, pitch, motivation


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

‘Music’ is both the observable product of intentional human action and a basic mode of thought by C9.P1
which any human action may be constituted. The most characteristic and e ective embodiment of
this mode of thought is what we call ‘music’ but it may also be manifested in other human
activities, and even in the organization of verbal ideas, such as Martin Luther King’s famous ‘I
have a dream’ speech or the poetry of Gertrude. (Blacking 1995: 224–225)

Time and measure are to instrumental Music what order and method are to discourse; they break it C9.P2
into proper parts and divisions, by which we are enabled both to remember better what has gone
before, and frequently to foresee somewhat of what is to come after: we frequently foresee the
return of a period which we know must correspond to another which we remember to have gone
before; and according to the saying of an ancient philosopher and musician, the enjoyment of
Music arises partly from memory and partly from foresight. (Smith 1980[1777/]: 204)
I think I have an idea of where I’m going and then I think […] for each tone you play, you could say C9.P3
it such that time was a long, long line and you had a lot of points on the line […] so for each point
you advance, then […] it gives […] it will give the premise for where the next point would be
because the tone itself in a sense de nes the next tone and so on because otherwise the phrase
becomes unnatural. And therefore you’re really in the tone you’re really on and then it kind of
gives itself how the next tone will become […] You need a smooth [development] and therefore
each tone, each and every vibrato, each and every oscillation, bowing, phrasing, everything builds

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
towards how it is going to become and it is impossible to predict how it will play out. It depends on
what you laid as ground.

(Fredrik Sjölin (cellist), quoted in Schiavio and Hø ding 2015: 13)

p. 150
Introduction C9.S1

SOMETHING we would describe as ‘music’ is widespread and ubiquitous across history and cultures. It is, C9.P4
overwhelmingly, an active integral part of human experience and reality. It is individually and collectively
embodied as well as being socially re ective and generative. Its meaning and signi cances necessarily
emerge and project in symbiosis with culture, the intricacies of interaction, and the individual experience
from which they are born (Cross 2003). Consequently, contrasting physical manifestations of music can
appear to reveal little, if any, de nite and observable common ground. Beyond an appreciation of basic
levels of surface structure, neither practical nor analytic musical skills translate across divides of cultural
style, immersion, and learning. Music can only ever be appropriately and fully heard, felt and understood
with a corresponding practice, knowledge, and experience of its complex histories, values, conventions,
institutions and technologies (see Cross 2001). Nevertheless, as something like music appears from our
cultural vantage point to be recognizable as such across cultures (Blacking 1995), it is logical to assume that
1
some identifying features or distinct conglomerations of features are commonly present. It is, furthermore,
consistently functional in common social settings and contexts. In countless and multifarious practices and
performances of music across time and geographical divides, we nd music to be functioning in regulating
emotional, cognitive, and physiological states; mediating between ‘self’ and ‘other’; representing cultural
symbolisms; and/or coordinating individual, dyadic, or group actions (see Clayton 2016; Nettl 2010). Any
notion, perhaps inspired by recent cultural and technological quirks, of music as something distinct,
pure/autonomous, or as a commercial product is highly problematic, and certainly fails to provide a
generalizable account of music and the shared roots of our human capacity for music (Bispham 2012). Our
experience with music and its situated e cacies is inextricably interwoven with the full mosaic of our
evolutionary past, our fundamental psychology and physiology; cultural and individual histories; most
central prosocial drives; and artistic motivations to express and create.

Crucial to the further course of argument in this chapter is the consideration that many, if not most, of C9.P5
music’s seemingly integral and most signi cant features are—perhaps extended or abstracted—but
certainly not unique to music. They are seeded in common predispositions and ontogenetic trajectories,
manifest in our earliest developmental explorations and are dynamically shaped throughout our lifespans.
Fundamentally in all physical movement and interaction we are continually creating rhythmic and quasi-
melodic shapes, sensing, engendering, and expressing our individual and collective vitalities (see Stern
2004; 2010). In terms, at least, of semantically describable categories of emotion, music seems to share with
vocal communication in general a common ‘code’ or set of cues (Juslin and Laukka 2003). Physical posture
and gesture too are inextricable parts in any fully manifest and successful communication (Kendon 2004),
p. 151 with joint actions and moments of synchronicity providing empathic attunements and generating points
of agreement and accord (Gill 2015). Mutually negotiated intrinsic motive pulses (Trevarthen 1999) and
tonal synchronies (Van Puyvelde et al. 2015) provide coordinative frameworks for interpersonal interaction,
intersubjectivity, attachment, participatory sense-making (De Jaegher and Di Paolo 2007), and the
regulation of a ect in parent–infant dyads and beyond (Stern 1985; 2009). In oratory and/or poetry,
dynamically generated expectancies and accents can add credence and persuasiveness (Woodall and
Burgoon 1981), while metre and cadence can allow emphasis, subvert expectations, highlight changes in
mood of interpretation, and/or aid in mnemonic retention. Additionally, broad analogies of
combinatoriality, embedded structure, implicitly acquired ‘grammars’, and recursion in music and
language show further overlap in terms both of macro design, organizational principles and of the cognitive

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
mechanisms of learning and working memory in particular that a ord them (see Rebuschat et al. 2011).

The full human capacity for music is clearly an intricate, interwoven, and phylogenetically emergent mosaic C9.P6
(see Foley 2012; 2016) overlapping in scope and relevance with all of the above. It is the result of a complex
evolutionary history of continuous and cumulative change. Each initially random mutation will have been
assembled gradually upon pre-existing biology and genetic coding, and will have endured for speci c
reasons and in response to particular environmental and social pressures. Regardless of continuing debates
concerning the speci cs, it is demonstrably a critical constituent part of our species and essential to a full
understanding of hominid evolution. Consequently, music is embedded in a erent and e erent connections
with more primary processes—our most foundational and central kinaesthetic, vocal, a ective, and socio-
intentional drives and capacities. Its many regulatory, a ective, and psychodynamic e ects and functions
2
are contingent upon the whole and the gregiological process of our species’ past. Most notably for the
current concerns in this book, a full account of music is necessarily contingent upon an appreciation of the
perception of time and the most central sensations of being in time. Even in trying to describe the basic
building blocks of a ‘simple’ musical pulse, for example, we are, from the outset, reliant upon describing
mechanisms that serve as reference points in the continuity of action in time (e.g. Thaut 2005; 2015) and
form our essential understanding of the ow of life events and phenomenological experience. The following
sections of this chapter will, therefore, o er a brief review of some of the more pertinent psychological
facets of our experience of time in relation not only to musical pulse but also to musical tone and musical
motivation. These will focus, in particular, on describing features that are particular to music, and will
propose that speci c features of the human capacity for music are most directly concerned with a ording a
form of sustained attention— a continuous linking between the immediate future, the perceptual present,
and future expectation. Although pulse may initially appear to have the most direct relevance to this
3
perspective and to the structuring of time in music, it is key to the argument presented here that the
intrinsic nature and speci cities of engaging with a musical pulse and musical tone and musical motivation
are all concerned essentially with crafting an extended experience in action and time—a musical moment.

p. 152
The Experience of Time C9.S2

Re ections on time and experience are, of course, long-standing and highly complex. Over 2000 years ago C9.P7
Aristotle deduced the continuity of time—its in nite divisibility—from the continuity of motion. This, in
turn, was deduced from the continuity of the space negotiated by any moving object. Time here is, as Gale
describes, ‘made continuous by the indivisible, present-now moment, which links the past to the future by
serving as the termination of the past and the beginning of the future’ (1968: 1). This is only the starting
point of intense philosophical debates, straddling our central understandings of physics and psychology,
which goes considerably beyond the current scope of discourse in this chapter.

From a psychological perspective we can understand Gale’s description of the experienced continuity of C9.P8
time, in conjunction with our fundamental nature, allowing us to perceive temporal ranges as being
simultaneous, sequential, owing/happening, present, experienced, and/or anticipated. Wittman and
Pöppel (1999), for example, describe three basic temporal experiences—simultaneity, non-simultaneity,
and temporal order—and have suggested that the temporal processing of sequential information can be
classi ed into four temporal ranges. Their description of the types of temporal order and the timespans
involved has stimulated some discussion and re nement (see Block and Gruber 2014). However, the
shortest two have been accepted as being pretty clear and uncontroversial (see Eisler et al. 2008): Between 0
and 2 ms, simultaneous and non-simultaneous events are perceived to be simultaneous (Hosokawa et al.
1981; Moore 1993), whereas events separated by 2/3 ms–20/40 ms create the impression of non-
simultaneity even though the temporal order of events cannot be con dently or unerringly distinguished
(Hirsh and Sherrick 1961; Lotze et al. 1999). In their review, Block and Gruber (2014) follow with timespans

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
of roughly 30–3,500 ms in which events appear as a changing present—there is an experience of time
owing and experience happening. This has been supported, for example, in paradigms of lm-
frames/snapshots being presented at increasing interstimulus intervals (e.g. men walking and bread being
toasted (Gruber and Block 2013)). Participants generally report that the ow of time starts to get lost at
about three seconds and has completely disappeared by seven seconds. In the same study, subjects were also
presented with the rst four notes of Beethoven’s fth symphony; participants no longer recognized the
opening phrase at an interstimulus interval above three seconds, and also reported that it did not sound like
music. This all ts well with earlier research arguing that three seconds is a time constant in perceptual
tasks representing a central neural mechanism that functions to integrate successive events into a ‘gestalt’
in order to create a ‘subjective present’ (Pöppel 1973; 1978). Pertinently for our concerns, it is within this
temporal range that repetitive isochronous events have been shown to easily build perceptual expectancies
towards future events (Jones 1976; Jones and Boltz 1989) and in which musical pulse is predominantly
p. 153 operational (see London 2012). Repp (2005: 973), for example, summarizes that the comfortable lower
and upper inter-onset-intervals limits in tapping tasks (roughly from 100–170 ms to 1,800–3,000 ms
(Bolton 1894; Fraisse 1982)) are musically relevant, in that they coincide approximately with ‘those within
which a sequence of events can be perceived to have rhythmic and metrical structure’.

Overlapping considerably with the three-second timespan is the notion of the ‘specious present’ (James C9.P9
1890/1950) or ‘living present’ (lebendige Gegenwart: Husserl 1991). In the simplest sense this is a subjective
experience of ‘now’—of the ‘present moment’. James’s original estimate for this phenomenon was that it
varied in length from a few seconds to ‘probably not more than minute’ (p. 642). Block and Gruber’s review
concludes that subsequent research suggests a time interval of about three seconds to, arguably, about seven
seconds, ‘during which the brain can compare and analyze very recent high-density memories in working
memory’ (p. 130). James famously describes that ‘the practically cognized present is no knife-edge, but a
saddle-back, with a certain breadth of its own on which we sit perched, and from which we look into two
directions into time’ (p. 609). Further on, he states that ‘its content is in a constant ux, events dawning
into its forward end as fast as they fade out of its rearward one […] meanwhile the specious present, the
intuited duration, stands permanent, like the rainbow on the waterfall, with its own quality unchanged by
the events that stream through it’ (p. 630). This idea that this ‘present’ enables our awareness of change
and sequence but is also itself a constant, invariable structural form of consciousness is also re ected in
Husserl’s characterization of the standing present (nunc stans). Husserl argues, however, that this is
‘ owing’ rather than static. He concludes: ‘what abides, above all, is the formal structure of the ow. That is
to say, the owing is not only owing throughout, but each phase has one and the same form’ (1991: 118). As
Dainton (2013: 391) summarizes, Husserl’s characterization ‘combines sameness—an ‘absolutely abiding
form’—with continual change and renewal […] it is an invariant structural feature of our consciousness’.

Although the question of whether the processing of larger timing intervals is operated by similar or C9.P10
overlapping mechanisms is open (Gibbon et al. 1997), it is widely accepted that durations exceeding three to
seven seconds involve an additional memory process that link moments that passed with the present
(Fraisse 1984). This further timing mechanism is thought to process the formation of perceptual gestalts by
binding successive events in perceptual units of two to three seconds (Pöppel 1978; 1997; also see Pöppel
4
and Wittmann 1999).
It is certainly no coincidence that music and melody constantly provide an ideal example in philosophical C9.P11
discourse on the binding of temporal experience (e.g. Husserl 1977). First and foremost, music is most
fundamentally a form of structured action in time and provides an almost perfect archetype of James’
famous ‘saddle-back’ quote. The signi cance of music’s inherent features depends holistically and
relationally on the shapes they create with what has come before and the further course that they inspire
and suggest. As testi ed by the three quotations at the head of this chapter, we are not, at the time of
musical immersion, primarily concerned with a mental representation of others, the communication of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
particular emotional states, or the complete acoustic outcome—the nal ‘portrait’. Instead, we are
p. 154 experiencing a moment—a psychological present that is constantly extended in relation to the timing,
tuning, and dynamic form of the immediate past and corresponding future-directed expectations and the
ongoing generation of action and attentional focus. To my mind this suggests a critical question. Does music
represent a paradigmatic example in our broad understanding of the experience of living in continuous time
as is implied in the philosophical discourse referenced above? Or is there something distinct about the
psychological mechanisms and phenomenological experience of time in music?

At rst glance, evidence for shared timing mechanisms across communicative domains, modalities, C9.P12
perception and performance (Ivry and Hazeltine 1995), and even across species (see Meck 2005), may
suggest the former. Although the precise nature of the central timing machinery—the ‘internal clock’—has
been the subject of a huge body of psychological modelling, experimental research, and debate (see Ivry and
Schlerf 2008; Grondin 2014), fundamental abilities to perceive and interact with the temporal structures of
the physical world and to direct our movements in time are certainly present in all higher-order senescent
animals. Leading researchers working on comparative timing between species con dently state, for
example, that ‘humans share with other animals an ability to measure the passage of physical time and
subjectively experience a sense of time passing’ (Allman et al. 2014), and that ‘an essential component of
primate cognitive function is the ability to extract and represent temporal information from the
environment. The quanti cation of the passage of time, in turn, is crucial to coordinate motor behavior’
5
(Zarco et al. 2009). Furthermore, although models of dynamic attending (Jones 1976) resonate strongly
with models of musical time (see Jones 2016) and o er a clear explanatory model for the psychology of
pulse in music (Large and Jones 1999), it is worth noting that the theory in itself seeks to explain our broad
capacities to interact with the temporalities of our movements and environments rather than music
speci cally. Once again music, with its commonly increased levels of predictability and stability, has
typically been presented as an ideal experimental testing ground and empirical paradigm for the theory,
rather than as something intrinsically separate.
Musical Pulse C9.S3

Evidence for a particular psychological manifestation of temporal structure in music does start to emerge in C9.P13
6
considering the distinction between relative/beat-based and absolute timing. Neuroscienti c and
neuropsychological studies seem to strongly suggest a di erence between beat-based and non-beat-based
timing. For example, in response to rhythms that induce a beat compared to those that do not, the basal

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
ganglia and supplementary motor areas are more active (Grahn and Brett 2007; Grahn and Rowe 2009).
Conversely, during absolute timing perception and motor tasks the cerebellum is more active (Grube et al.
2010; Teki et al. 2011). Further support for a functional distinction comes from fMRI studies of processing
p. 155 material that can be interpreted either relatively or in absolute terms. McAuley et al. (2006) devised an
interestingly ambiguous paradigmatic stimulus in which a periodic (600ms) beat is implied but not
explicitly emphasized. This leads to contrasting reports from participants of either slowing down or
speeding up at the end of a row of tones, thus indicating a di erence in whether the sequence was processed
in relation to the implied beat or in terms of more absolute interstimulus intervals. Crucially, both processes
were equally e ective: researchers found no evidence that duration-based timing was less accurate or
precise than beat-based timing. Subsequent neuroscienti c analysis (Grahn and McAuley 2009) supports
the idea that di erent strategies are supported by cortical activation di erences re ecting the engagement
7
of di erent neural timing mechanisms.

In active sensorimotor synchronization tasks (an experimental paradigm in which participants tap to match C9.P14
a range of metronomic and/or perturbed stimuli), we nd further evidence for some speci c capacity in
engaging with a ‘musical’ pulse (Repp 2005; Repp and Su 2013; Stevens 1886). In stark contrast to studies of
animal entrainment (Bispham 2006; 2018), this is a straightforward and absolutely universal skill in
humans. Even individuals categorized as being ‘amusic’ (Ayotte et al. 2002; Peretz 2006; Peretz et al. 2002)
typically show a normal ability to synchronize movement to the beat of popular dance music as well as
potential for improvement when given a modest amount of practice (Philips-Silver et al. 2013). In the one
reported case of ‘beat-deafness’, the individual in fact showed near-normal synchronization with a
metronome, suggesting that his de cit was concerned with nding the beat in music (Philips-Silver et al.
2011) rather than sensorimotor synchronization per se. Typically this ability is developed by the age of 4–6,
with age-speci c entrainment regions evidently narrower in childhood and late adulthood than in midlife
(see McAuley et al. 2006). ERP studies would appear to suggest further that the neurological foundations of
beat induction and corresponding future expectation are innately manifest (Winkler et al. 2009).

Another crucial distinction in the attempt to describe the particular psychological manifestation of temporal C9.P15
structure in music is between phase and period correction mechanisms. Sustaining regular sensorimotor
synchronization with other musicians or even with an arti cially isochronous pulse (such as a metronome)
requires some forms of corrective mechanisms. Without these, timing errors due to motor and/or internal
timekeeper variance (Wing and Kristo erson 1973) would simply accumulate and lead to a loss of synchrony
(Hary and Moore 1985). In tapping with another human or in a musical setting, we are constantly exploring
and reacting to further voluntary and involuntary uctuations due to features of individual style (Collier
1996; Collier and Collier 2002) as well as expressive and structurally motivated modulations of tempo and
microtimings (e.g. Iyer, 2002). Studies using experimentally controlled perturbations to the target stimuli
have provided a wealth of data and have been interpreted with regard to two principal models of error
correction mechanisms: dynamic systems theory (e.g. Schöner and Kelso 1988) and information-processing
theory (e.g. Vorberg and Wing 1996). Each case represents extensions of more general models of dynamic
attending (see above) or linear timekeeping (Wing and Beek 2002). Regardless of theoretical stance, two
p. 156 interacting yet behaviourally and neurologically distinct correction mechanisms are widely accepted to
be independently operational: phase correction and period correction (Mates 1994; Praamstra et al. 2003;
Repp 2001; Semjen et al. 1998; and also see Repp 2005). Phase correction essentially adjusts for
asynchronies between the last response and stimulus events assuming an unchanged period. Period
correction instead modi es the next target interval on the basis of discrepancies between the oscillatory or
timekeeper interval and the last or last few interstimulus intervals, thus altering the period of the
attentional musical pulse.

Crucially, phase correction seems to be a predominantly automatic process. It operates by and large without C9.P16
awareness in participants, and is equally e ective within and beyond perceptual thresholds for detecting

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
perturbations and/or asynchronies (Repp 2001; Repp and Penel 2002). Period correction is, however, a more
challenging cognitive task (e.g. Stephan et al. 2002). It is largely under cognitive control, requires
attentional resources, and relies on the conscious perception of a tempo change in the pacing sequence. This
was shown clearly in an important study by Repp and Keller (2004), who found that period correction is
dependent upon variables of intention, attentional load, and awareness, whereas phase correction was only
8
a ected partially by intention in supraliminal conditions. They therefore argue that phase correction and
period correction seem to represent independent processes of largely automatic action control and of
intentional cognitive control, respectively (Repp and Keller 2004). They hypothesize that their ndings
re ect the notion that ‘period correction is based on a more complex form of sensory evidence—namely on
a di erence between intervals (“relative period”)—than is phase correction, which is triggered by a
di erence between time points (or relative phase)’ (p. 517). In e ect they are highlighting the distinction
that period correction can be characterized as an interval comparison—a second-order di erence—
whereas phase correction results from a simple phase discrepancy—a rst-order di erence. The latter they
posit ‘even inanimate dynamic systems can perform’ (p. 517). Another way of explaining this idea, in the
context of the e ects of intention only, is given by Repp (2001), who hypothesized that the di erences were
due to ‘period correction requiring memory for at least one preceding event and as such greater
computational complexity calling for greater neural resources, thus making the process more extensive in
brain space and in time, and hence more accessible to higher-level cognitive processes’ (pp. 310–311).
Expanding on this in a later paper, Repp (2004: 76) states: ‘it is likely that period correction is a speci cally
human ability [and] is a manifestation of the more general human ability to set the tempo of a rhythmic
activity at will’ (see Bispham 2006; 2018).

Research and experimental ndings in sensorimotor synchronization are, of course, indicative of only some C9.P17
of the mechanisms involved in temporal coordination in music perception and performance. In musical
settings, innumerable additional social and interactional capabilities are operational in achieving temporal
and physical coordination. Interpersonal musical entrainment (Bispham 2003; Clayton et al. 2004) in social
contexts is, for example, inextricably embedded with our broad capacities for culture, action-mirroring, and
intersubjectivity (see Himberg 2014; Keller et al. 2014; Nowicki et al. 2013). A musical pulse is, in practice,
p. 157 in nitely more fascinating, variable, playful, and complex than the scope of the discussion and studies
considered above. It is manifest worldwide as an essentially boundless source of convention and creativity.
Nevertheless, we can, in my opinion, safely assume that these observations of two correction processes
re ect part of the basic temporal framework for musical rhythmic behaviours and interactions across
cultures. It even seems entirely plausible that the relative reliance on these foundational mechanisms in
music increases—revealed perhaps by a ‘stricter’ adherence to pulse—with the challenges of larger group
size and lack of social sympathy, familiarity, and experience. Broadly speaking, phase correction
mechanisms can be supposed to be operational in all activities involving future-directed attending where
expectations are constantly updated based upon asynchronies between attentional pulses and stimulus
events (Large and Jones 1999). In contrast, period correction is almost by de nition functional, speci cally,
within the framework of a sustained musical pulse. It is this latter process that arguably ties together an
awareness of the recent past, the psychological present, and our immediate expectancies into a longer
phenomenological experience. It also a ords impressive structures of group coordination and a collective
sense of shared time in music. I argue therefore that it forms part of a foundation for the speci cities of our
experience of time in music and some of the particular e cacies of music.
Musical Tone C9.S4

A similar argument to that set out here on musical pulse can be made for the most fundamental C9.P18
psychological architecture of a musical tone. The use of pitch in music, across composition, culture, and
time, is as immeasurably variable and fascinating as rhythm. It is ultimately a corresponding source of
individual character, cultural identity, playful exploration, socio-intentional drive, creativity, and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
emotional expression and regulation. I have suggested, however, that the capacity to engage with a ‘simple’
9
musical tone universally grounds the organizational use of pitch in music. Just as capacities to engage with
a basic musical pulse ground a wealth of complex rhythmic structure, the speci c nature of musical pitch
across cultures is built upon a primary ability to produce and engage with a sustained stable fundamental
frequency, and the ability to create or process certain relationships between pitches. Musical pitch structure
worldwide—whether monotone or orid—can be characterized as being relationally organized with
10
reference to sustained yet variable tonal areas (McAllester 1971). This applies, in principle, from the most
unadorned use of a persistent musical drone, through a private rendition of a folk melody, to instances of its
fullest potential across cultures in impressive group displays of choral harmony and instrumental
symphony. Crucially, however, this perspective does not necessarily imply any particular form of designed
tonal system or hierarchy. It merely suggests that—whether structured as sustained chanting, in relation to
a persistent underlying drone, as ‘chordal’ homophony, or as a complex polyphony of individual ‘voices’—
p. 158 musical pitch at any given time (or possibly within breath or phrase boundaries) is organized relationally
11
within a framework of a dominant pitch region or regions (Bispham 2009). In recent cross-cultural
research, for example, it has been shown that even in musical traditions featuring equitonic equal-spaced
scales (e.g. East African music) there is evidence to suggest that tonal centres are still perceived by idiomatic
12
listeners (Ross and Knight 2017).

The root structure of pitch in music is, therefore, in many ways similar to a sustained pulse. A musical tone C9.P19
is inherently a sustained physical and attentional practice that is rooted in particular correction
mechanisms (e.g. Natke et al. 2003; see also Larson and Robin 2016) and a degree of volitional control
(Keough et al. 2013). Initial experimental results further suggest that it is inextricably imbued with an
awareness of a pitch-structural framework (Hafke et al. 2013; 2016). Musical tone and musical pulse both
provide a mutually manifest focal point and framework for individual experience and social interaction. In
comparative analysis of both pulse and tone relational processing is a key and pronounced (if not absolute)
13
di erence between humans and other species (Hage and Nieder 2016; Patel and Demorest 2013). Most
notably for the principal concerns of this chapter, it is correction mechanisms that permit sustained
utterance and attention, a high degree of volitional control, and an awareness of architectural framework
that appear distinct to the context of music. Although the connection to time is less overt, it is in some sense
the same. Even in humming or listening to a single note, we are constantly integrating the events of the
previous moment with our current actions, and incessantly correcting and adapting our future
anticipations. Of further critical importance is that, in comparative analysis of tone and pulse, it is precisely
those features that distinguish the particular nature of individual experience with music that also a ord
group synchronous and harmonious interaction. E ectively, con gurations of musical pulse and musical
tone provide an attentional structure for managing personal experience in an extended perceptual present
—a continual phenomenological linking of the immediate past, the current moment (‘now’), and future
expectation—and a speci c architectural framework for interpersonal communication and an enacted sense
of shared time.
Musical Motivation C9.S5

To brie y summarize the course of discussion this far, not only can we posit something particular about the C9.P20
psychological mechanisms and phenomenological experience of time in music, but I contend that
investigating the speci cities of music’s essential architectural roots reveals that music is most
fundamentally distinct from other forms of communication in the way we structure attention and action in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
time. We can go further still and ask: what motivates us to put any given character or social action into
musical form? Why structure the ‘communicative musicalities’ of our dynamic mediations with others and
the ow of our individual experience around the speci cities of a musical pulse and/or musical tone? Here
again, I propose that the answer is directly related to the temporal structure of attention. As discussed
p. 159
brie y in the introduction, music can e ectively encapsulate and/or abstract all of our most central
vitalities, social drives, and cultural re ections (in varying combinations). It is intrinsically motivated by
our most central human motivations of emotional experience (Schiavio et al. 2017), expression (Juslin and
Sloboda 2011), and regulation (Gross and Thompson 2007); intersubjectivity (Trevarthen et al. 2011; Beebe
et al. 2005); shared intention (Tomasello et al. 2005); social alignment (Gallotti et al. 2017); cultural
belonging (Clayton 2003); and communitas (Rappaport 1999). It is a prominent component in our suite of
behaviours directed towards achieving hedonic and eudaimonic well-being through meaningful personal
re ection and socio-intentional connection; it is culturally generative and re ective, an artistic expression
of individual creativity, character, and cultural belonging. These motivations and their dynamic correlates
are, however, not speci cally enacted in music; they all emerge early in ontogeny, and are manifest in many
active and social forms in the course of human development. They are instead given and a orded particular
space in music—a space that constantly binds from one perceptual present to the next; o ers a constant
renewal of attentional focus; provides a degree of predictable continuity; and can be interactively manifest
in groups. Therefore, the distinct nature of a musical motivation lies in giving our central a ective and
socio-intentional drives extended phenomenological space, stability, and a degree of abstraction, intensity,
and meaning and in providing a framework for ritual action and for interpersonal and group interaction.

Musical Time C9.S6

The above discussion can be understood in terms of other recent work of mine on the speci c features of the C9.P21
human faculty for music—qualities that are at once universally present and operational in music across
cultures whilst also being speci c to our species and to the domain of music (see Bispham 2018; also termed
music’s evolutionary ‘design features’— Bispham 2009). In those papers, I conclude that unique features of
the human capacity for music—musical pulse, musical tone, and musical motivation—provide a sustained
attentional framework for managing personal experience and a coordinative structure for interpersonal
interaction. They are respectively embedded in and have developed from our most foundational and central
kinaesthetic: vocal, a ective, and socio-intentional drives and capacities. However, features that appear
unique to the context of musical engagement do not in or of themselves express or a ect anything. Rather,
they o er a sustainable socio-intentional coordinative strategy. They extend a phenomenological present,
thus a ording a continuing focus, and intensity of a ect, stability, regulation, memory, and meaning.
Accordingly, music can be understood primarily as an educational or therapeutic playground, a protracted,
liminal, and enhancing eld that allows a memorable event a ritualization and intensi cation of emotion
p. 160 and meaning. In accordance with Blacking’s (1995) description of it as a primary modelling system of
14
human thought, it is an indispensable tool of consciousness and a transformative technology of the mind.

Music, I suggest, should not be understood merely as a paradigmatic example of temporally dependent C9.P22
gestalt formation and our general experience of the continual ow of time. It is inextricably embedded with
our general timing mechanisms and fundamental capabilities to manage our environment and negotiate
interactions with others. However, it is also distinctly and profoundly a particular form of structuring
attention, action, and communicative intent. To be clear, I am not positing any kind of uniform experience.
Other chapters in this book, for example, testify to the varied nature of time perception across musical
styles and performances, and how it can be an integral and widespread source of compositional technique.
Another point to iterate clearly is that positing the speci cities of music and the corresponding experience
and attentional structure of time in its seemingly most basic rmament of a pulse, tone, and motivation
should not be understood as presenting a reductionist account of music, or as a position that discussion on

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
music or music in time should ever be exclusively framed in these terms. Rather, they constitute a
potentiating space, a particular experience of action in time that a ords and supports the extraordinary
exibility of dynamic form and function across cultures. They a ord an extended musical moment in which
a virtually in nite variety of individual creativity, emotional reality and regulation, social connection, group
coordination, and cultural history and re ection can be given a particular attentional shape. Therefore,
music’s distinct character lies in its being an ordered expression of human experience, behaviour,
interaction, and vitality, all shaped, shared, given signi cance, and/or transformed in time.

Notes
1. An important point to note is that these traits may not be physically or knowingly/descriptively evident but may reflect C9.N1
implicitly learned ontology and/or empathy with the underlying psychological mechanisms and experience.

2. Non-teleological. C9.N2

3. As testified e.g. by the predominant focus on pulse and rhythm in most chapters in this book. C9.N3

4. It is worth noting that these time-ranges seem to be operational across modalities, and in both perception and C9.N4
performance. Studies have shown e.g. that the threshold between events whose temporal order is or is not
distinguishable is consistent for acoustic, visual, and tactile stimuli (Hirsh and Sherrick 1961). Furthermore, timing
processes inherent in kinesthetic abilities is constant across a range of e ectors (Franz et al. 1992). Much empirical
evidence has, also, been collected to support the theoretical framework of the ʻpsychological presentʼ in perception
across modalities and in both perceptual and motor behaviour (see Wittmann and Pöppel 1999/2000).

5. In this model, which is less focused on judgements of time as on the operational side of how we and other animals C9.N5
interact in time with our environments, internal oscillatory mechanisms create points of attentional focus and expectation
p. 161 (Large and Jones 1999). Central to the theory is the view that attention is not a continuous operation but rather one in
which we constantly build expectancies and subsequently direct e icient energy pulses towards expected events whilst
reacting to the unexpected. A key feature of this model, therefore, is that it is attentionally future-directed with internal
predictions based upon perceived regularities and/or learned patterns of events entraining with the external world.

6. ʻRelative/beat-based timingʼ refers to the timing of intervals relative to a regular beat, whereas ʻabsolute timingʼ (also C9.N6
sometimes termed ʻduration-based timingʼ) describes the timing of absolute durations.

7. Relative timing appears also to have been a marked transition in human evolution. It seems that duration-based timing C9.N7
mechanisms are widely shared amongst primate species but that beat-based timing is either unique or particularly
advanced by many degrees in humans (e.g. Zarco et al. 2009; Donnet et al. 2014; Honing et al. 2012; Hattori et al. 2013;
Large and Gray 2015). Patel and Iverson (2014), Large and colleagues (2015), and Merchant and Honing (2014) o er three
contrasting (but possibly also complimentary) hypotheses on the evolution of beat-based timing in humans (see Bispham
2018 for discussion).

8. Correspondingly, di erent patterns of functional connectivity have also been shown depending on whether corrections C9.N8
are automatic or e ortful (Rao et al. 1997; Jäncke et al. 2000; Oullier et al. 2004; Chen et al. 2008; Thaut et al. 2009;
Bijsterbosch, Lee, Dyson-Sutton, et al. 2011; Bijsterbosch, Lee, Hunter, et al. 2011).

9. To clarify this should not be understood to imply that all use of pitch in music is necessarily structurally significant. Nor C9.N9
should it be interpreted to suggest that the specific ʻmusicnessʼ of a performance cannot reside entirely in the rhythmic
pulse. Instead I propose that music can be universally identified by engagement with configurations of musical pulse and
musical tone. One or the other could be entirely absent, but not both.

10. Although I am not aware of any direct experimental evidence it is interesting to consider that, in fast-moving or florid C9.N10
passages, individual notes are likely to be too fast to be the subject for correction of discrepancies between vocal or
instrumental production and an internal target or external reference (e.g. Lindblom and Sundberg 2007). Thus, we can
perhaps assume that these would be formed into a larger gestalt with the new whole constructed in reference to a more
sustained and internally manifest pitch centre. This, in turn, perhaps suggests a more realistic target for adjustment.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
11. The human ability to recognize and manifest pitch centre in complex musical structures is discussed in Podlipniak (2016). C9.N11
He argues, with reference to the ʻBaldwin e ectʼ, that this mental capacity emerged by joining the implicit recognition of
the frequency of pitch occurrence, working memory, and the emotional assessment of predicted stimuli.

12. Discrete tones have commonly been presented as one of the few distinguishing characteristics of music across cultures C9.N12
(Cross et al. 2013; Koelsch 2012; cf. Bispham 2009). More precisely and inclusively, however, these can be inferred as
resulting from, and being an expected acoustic correlate of, more deeply embedded processes. Essentially, a musical tone
is neither primarily defined by nor primarily experienced as absolute values or discrete scalar steps. It can be more or less
flexibly instantiated in relation to internal or externally influenced expectancies. More critical to its distinct nature than
any absolute acoustic attributes or analysis is that it is understood and psychologically maintained in reference to explicit
or implied pitch references or tonal centres. Similarly, rather than attempting to distinguish musical rhythm on the basis
of degrees of physical isochrony (Cross et al. 2013; Koelsch 2012; cf. Bispham 2009), a more primal way to consider the
specificities of a musical pulse is to consider the involvement of second-order relational processing over time, its inherent
p. 162 continuity, and correction mechanisms (see above). It is these features that profoundly create the stability of a musical
pulse. Relatively high degrees of behavioural isochrony are a particularly likely resultant attribute.

13. Most likely this reflects and follows a wider move towards increased sociality in humans and a motivation to understand C9.N13
others as ourselves—in terms empathetically of our own psychological and physical experience.

14. I am grateful to an anonymous reviewer who notes that the conception of musical structure developed in this chapter C9.N14
could also function as a model or architecture for other human thought. It could, he/she suggests, be a reflection, an
artistic mimesis, of the structure of human thought prior to music. This is certainly an interesting idea that could be
fruitfully explored further.
References C9.S7

Allman, M. J., Teki, S., Gri iths, T. D., and Meck, W. H. (2014). Properties of the internal clock: first-and second-order principles of C9.P23
subjective time. Annual Review of Psychology 65(1): 743–771.
Google Scholar WorldCat

Ayotte, J., Peretz, I., and Hyde, K. (2002). Congenital amusia: a group study of adults a licted with a music‐specific disorder. C9.P24

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Brain 125(2): 238–251.
Google Scholar WorldCat

Beebe, B., Knoblauch, S., Rustin, J., and Sorter, D. (2005). Forms of intersubjectivity in infant research and adult treatment. Other C9.P25
Press.
Google Scholar Google Preview WorldCat COPAC

Bijsterbosch, J. D., Lee, K.-H., Dyson-Sutton, W., Barker, A. T., and Woodru , P. W. R. (2011). Continuous theta burst stimulation C9.P26
over the le pre-motor cortex a ects sensorimotor timing accuracy and supraliminal error correction. Brain Research 1410: 101–
111.
Google Scholar WorldCat

Bijsterbosch, J. D., Lee, K.-H., Hunter, M.D., Tsio, D. T., Lankappa, S., Wildinson, I. D., and Woodru , P. W. R. (2011). The role of the C9.P27
cerebellum in sub-and supraliminal error correction during sensorimotor synchronization: evidence from fMRI and TMS. Journal
of Cognitive Neuroscience 23(5): 1100–1112.
Google Scholar WorldCat

Bispham, J. (2003). An evolutionary perspective on the human skill of interpersonal musical entrainment (masterʼs thesis). C9.P28
University of Cambridge.
Google Scholar Google Preview WorldCat COPAC

Bispham, J. (2006). Rhythm in music: What is it? Who has it? And why? Music Perception 24(2): 125–134. C9.P29
Google Scholar WorldCat

Bispham, J. (2009). Musicʼs ʻdesign featuresʼ: Musical motivation, musical pulse, and musical pitch. Musicae Scientiae 13(2) C9.P30
(suppl.): 41–61.
Google Scholar WorldCat

Bispham, J. (2012). How musical is man? An evolutionary perspective. In A. R. Brown (ed.), Sound musicianship: understanding C9.P31
the cra s of music, 126–137. Cambridge Scholars.
Google Scholar Google Preview WorldCat COPAC

Bispham, J. (2018). The human capacity for music: Whatʼs special about it? (doctoral dissertation). University of Cambridge. C9.P32
https://doi.org/10.17863/CAM.31835
Google Scholar Google Preview WorldCat COPAC

Blacking, J. (1995). Music, culture, and experience: Selected papers of John Blacking. University of Chicago Press. C9.P33
Google Scholar Google Preview WorldCat COPAC

Block, R. A., and Gruber, R. P. (2014). Time perception, attention, and memory: a selective review. Acta Psychologica 149: 129– C9.P34
133.
Google Scholar WorldCat

Bolton, T. L. (1894). Rhythm. American Journal of Psychology 6(2): 145–238. C9.P35


Google Scholar WorldCat

Chen, J. L., Penhune, V. B., and Zatorre, R. J. (2008). Moving on time: brain network for auditory–motor synchronization is C9.P36
modulated by rhythm complexity and musical training. Journal of Cognitive Neuroscience 20(2): 226–239.
Google Scholar WorldCat

p. 163 Clayton, M. (2016). The social and personal functions of music in cross-cultural perspective. In S. Hallam, I. Cross, and M. Thaut C9.P37
(eds), The Oxford handbook of music psychology (2nd edn), 35–44. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Clayton, M., Herbert, T., and Middleton, R. (eds) (2003). The cultural study of music: A critical introduction. Routledge. C9.P38

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Clayton, M., Sager, R., and Will, U. (2004). In time with the music: The concept of entrainment and its significance for C9.P39
ethnomusicology. ESEM Counterpoint 11: 3–142.
Google Scholar WorldCat

Collier, G. L. (1996). Microrhythms in jazz: a review of papers. Annual Review of Jazz Studies 8: 117–139. C9.P40
Google Scholar WorldCat

Collier, G. L., and Collier, J. L. (2002). A study of timing in two Louis Armstrong solos. Music Perception 19(3): 463–483. C9.P41
Google Scholar WorldCat

Cross, I. (2001). Music, cognition, culture, and evolution. Annals of the New York Academy of Sciences 930(1): 28–42. C9.P42
Google Scholar WorldCat

Cross, I. (2003). Music and biocultural evolution. In M. Clayton, T. Herbert, and R. Middleton (eds), The cultural study of music: A C9.P43
critical introduction, 19–30. Routledge.
Google Scholar Google Preview WorldCat COPAC

Cross, I., Fitch, W. T., Aboitiz, F., Iriki, A., Jarvis, E. D., Lewis, J., Liebal, K., Merker, B., Stout, D., and Trehub, S. E. (2013). Culture C9.P44
and evolution. In Michael Arbib (ed.), Language, music and the brain, 541–562. MIT Press.
Google Scholar Google Preview WorldCat COPAC

Dainton, B. (2013). The perception of time. In A. Bardon, and H. Dyke (eds), A companion to the philosophy of time, 389–409. C9.P45
Wiley.
Google Scholar Google Preview WorldCat COPAC

De Jaegher, H., and Di Paolo, E. (2007). Participatory sense-making. Phenomenology and the Cognitive Sciences 6(4): 485–507. C9.P46
Google Scholar WorldCat

Donnet, S., Bartolo, R., Fernandes, J. M., Silva Cunha, J. P., Prado, L., and Merchant, H. (2014). Monkeys time their pauses of C9.P47
movement and not their movement-kinematics during a synchronization-continuation rhythmic task. Journal of
Neurophysiology 111(10): 2138–2149.
Google Scholar WorldCat

Eisler, H., Eisler, A. D., and Hellström, A. (2008). Psychophysical issues in the study of time perception. In S. Grondin (ed.), C9.P48
Psychology of time, 75–109. Emerald.
Google Scholar Google Preview WorldCat COPAC

Foley, R. A. (2012). Music and mosaics: The evolution of human abilities. In N. Bannan (ed.), Music, language, and human C9.P49
evolution, 31–57. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Foley, R. A. (2016). Mosaic evolution and the pattern of transitions in the hominin lineage. Philosophical Transactions of the Royal C9.P50
Society B: Biological Sciences 371. http://dx.doi.org/10.1098/rstb.2015.0244
Google Scholar WorldCat

Fraisse, P. (1982). Rhythm and tempo. In D. Deutsch (ed.), The psychology of music, 149–180. Elsevier Science. C9.P51
Google Scholar Google Preview WorldCat COPAC

Fraisse, P. (1984). Perception and estimation of time. Annual Review of Psychology 35(1): 1–37. C9.P52
Google Scholar WorldCat

Franz, E. A., Zelaznik, H. N., and Smith, A. (1992). Evidence of common timing processes in the control of manual, orofacial, and C9.P53
speech movements. Journal of Motor Behavior 24(3): 281–287.
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Gale, R. (1968). The philosophy of time: A collection of essays. Palgrave Macmillan. C9.P54
Google Scholar Google Preview WorldCat COPAC

Gallotti, M., Fairhurst, M., and Frith, C. (2017). Alignment in social interactions. Consciousness and Cognition 48: 253–261. C9.P55
https://doi:10.1016/j.concog.2016.12.002
Google Scholar WorldCat

Gibbon, J., Malapani, C., Dale, C. L, and Gallistel, C. (1997). Toward a neurobiology of temporal cognition: advances and C9.P56
challenges. Current Opinion in Neurobiology 7(2): 170–184.
Google Scholar WorldCat

Gill, S. P. (2015). Tacit engagement. In Tacit engagement: Beyond interaction, 1–34. Springer. C9.P57

Grahn, J. A., and Brett, M. (2007). Rhythm and beat perception in motor areas of the brain. Journal of Cognitive Neuroscience C9.P58
19(5): 893–906.
Google Scholar WorldCat

p. 164 Grahn, J. A., and McAuley, J. D. (2009). Neural bases of individual di erences in beat perception. Neuroimage 47(4): 1894–1903. C9.P59
Google Scholar WorldCat

Grahn, J. A., and Rowe, J. B. (2009). Feeling the beat: premotor and striatal interactions in musicians and non-musicians during C9.P60
beat perception. Journal of Neuroscience 29(23): 7540–7548.
Google Scholar WorldCat

Grondin, S. (2014). About the (non) scalar property for time perception. In H. Merchant and V. de Lafuente (eds), Neurobiology of C9.P61
interval timing, 17–32. Springer.
Google Scholar Google Preview WorldCat COPAC

Gross, J. J., and Thompson, R. A. (2007). Emotion regulation: Conceptual foundations. In J. J. Gross (ed.), Handbook of emotion C9.P62
regulation, 3–24. Guilford Press.
Google Scholar Google Preview WorldCat COPAC

Grube, M., Lee, K.-H., Gri iths, T. D., Barker, A. T., and Woodru , W. R. (2010). Transcranial magnetic theta-burst stimulation of the C9.P63
human cerebellum distinguishes absolute, duration-based from relative, beat-based perception of subsecond time intervals.
Frontiers in Psychology 1(171): 1–8. https://doi.org/10.3389/fpsyg.2010.00171
Google Scholar WorldCat

Gruber, R. P., and Block, R. A. (2013). The flow of time as a perceptual illusion. Journal of Mind and Behavior 34(1): 91–100. C9.P64
Google Scholar WorldCat

Hafke-Dys, H. Z., Preis, A., and Kaczmarek, T. (2013). Comparison of perceptual and motor responses to changes in intensity and C9.P65
voice fundamental frequency. Acta Acustica united with Acustica 99(3): 457–464.
Google Scholar WorldCat

Hafke-Dys, H., Preis, A., and Trojan, D. (2016). Violinistsʼ perceptions of and motor reactions to fundamental frequency shi s C9.P66
introduced in auditory feedback. Acta Acustica united with Acustica 102(1): 155–158.
Google Scholar WorldCat

Hage, S. R., and Nieder, A. (2016). Dual neural network model for the evolution of speech and language. Trends in Neurosciences C9.P67
39(12): 813–829.
Google Scholar WorldCat

Hary, D., and Moore, G. P. (1985). Temporal tracking and synchronization strategies. Human Neurobiology 4(2): 73–79. C9.P68
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Hattori, Y., Tomonaga, M., and Matsuzawa, T. (2013). Spontaneous synchronized tapping to an auditory rhythm in a chimpanzee. C9.P69
Scientific Reports 3. https://doi.org/10.1038/srep01566
Google Scholar WorldCat

Himberg, T. (2014). Interaction in musical time (doctoral dissertation). University of Cambridge. C9.P70
Google Scholar Google Preview WorldCat COPAC

Hirsh, I. J., and Sherrick, C. E. Jr (1961). Perceived order in di erent sense modalities. Journal of Experimental Psychology 62(5): C9.P71
423–432.
Google Scholar WorldCat

Honing, H., Merchant, H. Haden, G. P., Prado, L., and Bartolo, R. (2012). Rhesus monkeys (Macaca mulatta) detect rhythmic C9.P72
groups in music, but not the beat. PloS ONE 7(12): e51369. https://doi.org/10.1371/journal.pone.0051369
Google Scholar WorldCat

Hosokawa, T., Nakamura, R., and Shibuya, N. (1981). Monotic and dichotic fusion thresholds in patients with unilateral C9.P73
subcortical lesions. Neuropsychologia 19(2): 241–248.
Google Scholar WorldCat

Husserl E. (1977). Phenomenological psychology: Lectures, summer semester (1925) (trans. J. Scanlon). Martinus Nijho . C9.P74
Google Scholar Google Preview WorldCat COPAC

Husserl, E. (1991). On the phenomenology of the consciousness of internal time. Kluwer Academic. C9.P75
Google Scholar Google Preview WorldCat COPAC

Ivry, R. B., and Hazeltine, R. E. (1995). Perception and production of temporal intervals across a range of durations: evidence for a C9.P76
common timing mechanism. Journal of Experimental Psychology: Human Perception and Performance 21(1): 3–18.
Google Scholar WorldCat

Ivry, R. B., and Schlerf, J. E. (2008). Dedicated and intrinsic models of time perception. Trends in Cognitive Sciences 12(7): 273– C9.P77
280.
Google Scholar WorldCat

Iyer, V. (2002). Embodied mind, situated cognition, and expressive microtiming in African-American music. Music Perception C9.P78
19(3): 387–414.
Google Scholar WorldCat

p. 165 James, W. (1950[1890]). The principles of psychology. Dover. C9.P79


Google Scholar Google Preview WorldCat COPAC

Jäncke, L., Loose, R., Lutz, K., Specht, K., and Shah, N. J. (2000). Cortical activations during paced finger-tapping applying visual C9.P80
and auditory pacing stimuli. Cognitive Brain Research 10(1): 51–66.
Google Scholar WorldCat

Jones, M. R. (1976). Time, our lost dimension: Toward a new theory of perception, attention, and memory. Psychological Review C9.P81
83(5): 325–335.
Google Scholar WorldCat

Jones, M. R. (2016). Musical time. In S. Hallam, I. Cross, and M. Thaut (eds), The Oxford handbook of music psychology, 125–141. C9.P82
Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Jones, M. R., and Boltz, M. (1989). Dynamic attending and responses to time. Psychological Review 96(3): 459–491. C9.P83
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Juslin, P., and Laukka, P. (2003). Communication of emotions in vocal expression and music performance: Di erent channels, C9.P84
same code? Psychological Bulletin 129(5): 770–814.
Google Scholar WorldCat

Juslin, P. N., and Sloboda, J. (2011). Handbook of music and emotion: Theory, research, applications. Oxford University Press. C9.P85
Google Scholar Google Preview WorldCat COPAC

Keller, P. E., Novembre, G., and Hove, M. J. (2014). Rhythm in joint action: psychological and neurophysiological mechanisms for C9.P86
real-time interpersonal coordination. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences
369(1658). https://doi.org/10.1098/rstb.2013.0394
Google Scholar WorldCat

Kendon, A. (2004). Gesture: Visible action as utterance. Cambridge University Press. C9.P87
Google Scholar Google Preview WorldCat COPAC

Keough, D., Hawco, C., and Jones, J. A. (2013). Auditory–motor adaptation to frequency-altered auditory feedback occurs when C9.P88
participants ignore feedback. BMC Neuroscience 14: 25. https://doi.org/10.1186/1471-2202-14-25
Google Scholar WorldCat

Koelsch, S. (2012). Brain and music. Wiley. C9.P89


Google Scholar Google Preview WorldCat COPAC

Large, E. W., and Gray, P. M. (2015). Spontaneous tempo and rhythmic entrainment in a bonobo (Pan paniscus). Journal of C9.P90
Comparative Psychology 129(4): 317–328. https://doi.org/10.1037/com0000011
Google Scholar WorldCat

Large, E. W., Herrera, J. A., and Velasco, M. J. (2015). Neural networks for beat perception in musical rhythm. Frontiers in Systems C9.P91
Neuroscience 9(159). https://doi:10.3389/fnsys.2015.0015
Google Scholar WorldCat

Large, E. W., and Jones, M. R. (1999). The dynamics of attending: How people track time–varying events. Psychological Review C9.P92
106(1): 119–159.
Google Scholar WorldCat

Larson, C. R., and Robin, D. A. (2016). Sensory processing: Advances in understanding structure and function of pitch-shi ed C9.P93
auditory feedback in voice control. AIMS Neuroscience 3(1): 22–39. https://doi:10.3934/Neuroscience.2016.1.22
Google Scholar WorldCat

Lindblom, B., and Sundberg, J. (2007). The human voice in speech and singing. In T. D. Rossing (ed.), Springer handbook of C9.P94
acoustics, 669–712. Springer Science.
Google Scholar Google Preview WorldCat COPAC

London, J. (2012). Hearing in time: Psychological aspects of musical metre (2nd edn). Oxford University Press. C9.P95
Google Scholar Google Preview WorldCat COPAC

Lotze, M., Wittmann, M., von Steinbüchel, N., Pöppel, E., and Roenneberg, T. (1999). Daily rhythm of temporal resolution in the C9.P96
auditory system. Cortex 35(1): 89–100. https://doi.org/10.1016/S0010-9452(08)70787-1
Google Scholar WorldCat

Mates, J., Müller, U., Radil, T., and Pöppel, E. (1994). Temporal integration in sensorimotor synchronization. Journal of Cognitive C9.P97
Neuroscience 6(4): 332–340.
Google Scholar WorldCat

McAllester, D. P. (1971). Some thoughts on ʻuniversalsʼ in world music. Ethnomusicology 15(3): 379–380. C9.P98

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

McAuley, J. D., Jones, M. R., Holub, S., Johnston, H. M., and Miller, N. S. (2006). The time of our lives: Life span development of C9.P99
timing and event tracking. Journal of Experimental Psychology: General 135(3): 348–367. https://doi.org/10.1037/0096-
3445.135.3.348
Google Scholar WorldCat

p. 166 Meck, W. H. (2005). Neuropsychology of timing and time perception. Brain and Cognition 58(1): 1–8. C9.P100
Google Scholar WorldCat

Merchant, H., and Honing, H. (2014). Are non-human primates capable of rhythmic entrainment? Evidence for the gradual C9.P101
audiomotor evolution hypothesis. Frontiers in Neuroscience 7: 274. https://doi.org/10.3389/fnins.2013.00274
Google Scholar WorldCat

Moore, B. C. (1993). Temporal analysis in normal and impaired hearing. Annals of the New York Academy of Sciences 682(1): 119– C9.P102
136.
Google Scholar WorldCat

Natke, U., Donath, T. M., and Kalveram, K. T. (2003). Control of voice fundamental frequency in speaking versus singing. Journal C9.P103
of the Acoustical Society of America 113(3): 1587–1593. https://doi.org/10.1121/1.1543928
Google Scholar WorldCat

Nettl, B. (2010). The study of ethnomusicology: Thirty-one issues and concepts (3rd edn). University of Illinois Press. C9.P104
Google Scholar Google Preview WorldCat COPAC

Nowicki, L., Prinz, W., Grosjean, M., Repp, B. H., and Keller, P. E. (2013). Mutual adaptive timing in interpersonal action C9.P105
coordination. Psychomusicology: Music, Mind, and Brain 23(1): 6–20. https://doi.org/10.1037/a0032039
Google Scholar WorldCat

Oullier, O., Jantzen, K. J., Steinberg, F. L., and Kelso, J. A. (2004). Neural substrates of real and imagined sensorimotor C9.P106
coordination. Cerebral Cortex 15(7): 975–985.
Google Scholar WorldCat

Patel, A., and Demorest, S. (2013). Comparative music cognition: cross-species and cross-cultural studies. In D. Deutsch (ed.), The C9.P107
psychology of music (3rd edn), 647–681. Elsevier Science.
Google Scholar Google Preview WorldCat COPAC

Patel, A. D., and Iversen, J. R. (2014). The evolutionary neuroscience of musical beat perception: the Action Simulation for C9.P108
Auditory Prediction (ASAP) hypothesis. Frontiers in Systems Neuroscience 8: 57. https://doi.org/10.3389/fnsys.2014.00057
Google Scholar WorldCat

Peretz, I. (2006). The nature of music from a biological perspective. Cognition 100(1): 1–32. C9.P109
Google Scholar WorldCat

Peretz, I., Ayotte, J., Zatorre, R. J., Mehler, J., Ahad, P., Penhune, V. B., and Jutras, B. (2002). Congenital amusia: a disorder of fine- C9.P110
grained pitch discrimination. Neuron 33(2): 185–191. https://doi.org/10.1016/s0896-6273(01)00580-3
Google Scholar WorldCat
Phillips-Silver, J., Toiviainen, P., Gosselin, N., and Peretz, I. (2013). Amusic does not mean unmusical: Beat perception and C9.P111
synchronization ability despite pitch deafness. Cognitive Neuropsychology 30(5): 311–331.
Google Scholar WorldCat

Phillips-Silver, J., Toiviainen, P., Gosselin, N., Piché, O., Nozaradan, S., Palmer, C., and Peretz, I. (2011). Born to dance but beat C9.P112
deaf: a new form of congenital amusia. Neuropsychologia 49(5): 961–969.
https://doi.org/10.1016/j.neuropsychologia.2011.02.002
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Podlipniak, P. (2016). The evolutionary origin of pitch centre recognition. Psychology of Music 44(3): 527–543. C9.P113
Google Scholar WorldCat

Pöppel, E. (1973). Influence of pause duration on the reproduction of a 2-second interval. Bulletin of the Psychonomic Society C9.P114
2(5): 291–292.
Google Scholar WorldCat

Pöppel E. (1978). Time perception. In R. Held, H. W. Leibowitz, and H. L. Teuber (eds), Perception: Handbook of sensory C9.P115
physiology, vol. 8, 713–729. Springer. https://doi.org/10.1007/978-3-642-46354-9_23
Google Scholar Google Preview WorldCat COPAC

Pöppel, E. (1997). A hierarchical model of temporal perception. Trends in Cognitive Sciences 1(2): 56–61. C9.P116
Google Scholar WorldCat

Pöppel, E., and Wittmann, M. (1999). Time in the mind. In R. Wilson and F. Keil (eds), The MIT encyclopedia of the cognitive C9.P117
sciences, 841–843. MIT Press.
Google Scholar Google Preview WorldCat COPAC

Praamstra, P., Turgeon, M., Hesse, C. W., Wing, A. M., and Perryer, L. (2003). Neurophysiological correlates of error correction in C9.P118
sensorimotor-synchronization. Neuroimage 20(2): 1283–1297.
Google Scholar WorldCat

p. 167 Rao, S. M., Harrington, D. L., Haaland, K. Y., Bobholz, J. A., Cox, R. W., and Binder, J. R. (1997). Distributed neural systems C9.P119
underlying the timing of movements. Journal of Neuroscience 17(14): 5528–5535.
Google Scholar WorldCat

Rappaport, R. A. (1999). Ritual and religion in the making of humanity. Cambridge University Press. C9.P120
Google Scholar Google Preview WorldCat COPAC

Rebuschat, P., Rohrmeier, M., Hawkins, J. A., and Cross, I. (2011). Language and music as cognitive systems. Oxford University C9.P121
Press.
Google Scholar Google Preview WorldCat COPAC

Repp, B. H. (2001). Processes underlying adaptation to tempo changes in sensorimotor synchronization. Human Movement C9.P122
Science 20(3): 277–312.
Google Scholar WorldCat

Repp, B. (2004). Comments on ʻRapid motor adaptations to subliminal frequency shi s during syncopated rhythmic C9.P123
sensorimotor synchronisationʼ by M. Thaut and G. Kenyon. Human Movement Science 21(3): 61–78.
Google Scholar WorldCat

Repp, B. H. (2005). Sensorimotor synchronization: a review of the tapping literature. Psychonomic bulletin and review 12(6): 969– C9.P124
992.
Google Scholar WorldCat

Repp, B. H., and Keller, P. E. (2004). Adaptation to tempo changes in sensorimotor synchronization: E ects of intention, attention, C9.P125
and awareness. Quarterly Journal of Experimental Psychology Section A, 57(3): 499–521.
Google Scholar WorldCat

Repp, B. H., and Penel, A. (2002). Auditory dominance in temporal processing: new evidence from synchronization with C9.P126
simultaneous visual and auditory sequences. Journal of Experimental Psychology: Human Perception and Performance 28(5):
1085–1099.
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Repp, B. H., and Su, Y.–H. (2013). Sensorimotor synchronization: a review of recent research (2006–2012). Psychonomic Bulletin C9.P127
and Review 20(3): 403–452.
Google Scholar WorldCat

Ross, B., and Knight, S. (2017). Reports of equitonic scale systems in African musical traditions and their implications for C9.P128
cognitive models of pitch organization. Musicae Scientiae 13(2): 231–272.
Google Scholar WorldCat

Schiavio, A., and Hø ding, S. (2015). Playing together without communicating? A pre–reflective and enactive account of joint C9.P129
musical performance. Musicae Scientiae 19(4): 366–388.
Google Scholar WorldCat

Schiavio, A., van der Schy , D., Cespedes-Guevara, J., and Reybrouck, M. (2017). Enacting musical emotions: Sense-making, C9.P130
dynamic systems, and the embodied mind. Phenomenology and the Cognitive Sciences 16(5): 785–809.
https://doi.org/10.1007/s11097-016-9477-8
Google Scholar WorldCat

Schöner, G., and Kelso, J. (1988). A synergetic theory of environmentally-specified and learned patterns of movement C9.P131
coordination. Biological Cybernetics 58(2): 71–80.
Google Scholar WorldCat

Semjen, A., Vorberg, D., and Schulze, H.-H. (1998). Getting synchronized with the metronome: Comparisons between phase and C9.P132
period correction. Psychological Research 61(1): 44–55. https://doi.org/10.1007/s004260050012
Google Scholar WorldCat

Smith, A. (1777/1980). Of the nature of that imitation which takes place in what are called the imitative arts. Essays on C9.P133
philosophical subjects. In W. P. D. Wightman, J. C. Bryce, and I. S. Ross (eds), The Glasgow edition of the works and
correspondence of Adam Smith, vol. 3: Essays on philosophical subjects with Dugald Stewartʼs account of Adam Smith, 176–213.
Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Stephan, K. M., Thaut, M. H., Wunderlich, G., Schicks, W., Tian, B., Tellmann, L., Schmitz, T., Herzog, H., McIntosh, G. C., Xietz, R. J., C9.P134
and Hömberg, V. (2002). Conscious and subconscious sensorimotor synchronization: Prefrontal cortex and the influence of
awareness. Neuroimage 15(2): 345–352.
Google Scholar WorldCat

Stern, D. N. (1985). The interpersonal world of the infant: A view from psychoanalysis and developmental psychology. Basic Books. C9.P135
Google Scholar Google Preview WorldCat COPAC

p. 168 Stern, D. N. (2004). The present moment in psychotherapy and everyday life. Norton. C9.P136
Google Scholar Google Preview WorldCat COPAC

Stern, D. N. (2009). The first relationship. Cambridge, MA: Harvard University Press. C9.P137
Google Scholar Google Preview WorldCat COPAC

Stern, D. N. (2010). Forms of vitality: Exploring dynamic experience in psychology, the arts, psychotherapy, and development. C9.P138
Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Stevens, L. T. (1886). On the time-sense. Mind 11(43): 393–404. C9.P139


Google Scholar WorldCat

Teki, S., Grube, M., Kumar, S., and Gri iths, T. D. (2011). Distinct neural substrates of duration-based and beat-based auditory C9.P140
timing. Journal of Neuroscience 31(10): 3805–3812.
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Thaut, M. H. (2005). Rhythm, music, and the brain: Scientific foundations and clinical applications. Routledge. C9.P141
Google Scholar Google Preview WorldCat COPAC

Thaut, M. H. (2015). The discovery of human auditory–motor entrainment and its role in the development of neurologic music C9.P142
therapy. Progress in Brain Research 217: 253–266.
Google Scholar WorldCat

Thaut, M. H., Stephan, K. M., Wunderlich, G., Schicks, W., Tellmann, L., Herzog, H., McIntosh, G. C., Seitz, R. J., and Hömberg, V. C9.P143
(2009). Distinct cortico-cerebellar activations in rhythmic auditory motor synchronization. Cortex 45(1): 44–53.
Google Scholar WorldCat

Tomasello, M., Carpenter, M., Call, J., Behne, T., and Moll, H. (2005). Understanding and sharing intentions: The origins of cultural C9.P144
cognition. Behavioral and Brain Sciences 28(5): 675–691.
Google Scholar WorldCat

Trevarthen, C. (1999). Musicality and the intrinsic motive pulse: evidence from human psychobiology and infant communication. C9.P145
Musicae Scientiae 3 (suppl.): 155–215.
Google Scholar WorldCat

Trevarthen, C., Delafield–Butt, J., and Schögler, B. (2011). Psychobiology of musical gesture: innate rhythm, harmony and C9.P146
melody. In A. Gritten, and E. King (eds), New perspectives on music and gesture, 11–43. Routledge.
Google Scholar Google Preview WorldCat COPAC

Van Puyvelde, M., Loots, G., Gillisjans, L., Pattyn, N., and Quintana, C. (2015). A cross-cultural comparison of tonal synchrony and C9.P147
pitch imitation in the vocal dialogs of Belgian Flemish-speaking and Mexican Spanish-speaking mother–infant dyads. Infant
Behavior and Development 40: 41–53.
Google Scholar WorldCat

Vorberg, D., and Wing, A. (1996). Modeling variability and dependence in timing. In H. Heuer and S. W. Keele (eds), Handbook of C9.P148
perception and action, vol. 2, 181–262. Academic Press.
Google Scholar Google Preview WorldCat COPAC

Wing, A. M., and Beek, P. J. (2002). Movement timing: a tutorial. In W. Prinz and B. Hommel (eds), Common mechanisms in C9.P149
perception and action: Attention and performance, vol. 19, 202–226. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Wing, A. M., and Kristo erson, A. (1973). The timing of interresponse intervals. Attention, Perception, and Psychophysics 13(3): C9.P150
455–460.
Google Scholar WorldCat

Winkler, I., Háden, G. P., Ladinig, O., Sziller, I., and Honing, H. (2009). Newborn infants detect the beat in music. Proceedings of C9.P151
the National Academy of Sciences 106(7): 2468–2471. https://doi.org/10.1073/pnas.0809035106
Google Scholar WorldCat

Wittmann, M., and Pöppel, E. (1999). Temporal mechanisms of the brain as fundamentals of communication—with special C9.P152
reference to music perception and performance. Musicae Scientiae 3(suppl.): 13–28.
Google Scholar WorldCat

Woodall, W. G., and Burgoon, J. K. (1981). The e ects of nonverbal synchrony on message comprehension and persuasiveness. C9.P153
Journal of Nonverbal Behavior 5(4): 207–223.
Google Scholar WorldCat

Zarco, W., Merchant, H., Prado, L., and Mendez, J. C. (2009). Subsecond timing in primates: comparison of interval production C9.P154
between human subjects and rhesus monkeys. Journal of Neurophysiology 102(6): 3191–3202.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469499 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
CHAPTER

10 Timescales and the Temporal Emergence of Musicking  C10

Juan M. Loaiza

https://doi.org/10.1093/oxfordhb/9780190947279.013.15 Pages 169–196


Published: 08 December 2021

Abstract
The aim of this chapter is twofold: to present a new way of mapping timescales of musicking, and to
elaborate an explanatory approach that overcomes philosophical reductionism and allows
interdisciplinary conversation. It proposes that the emergence of organizational properties in
musicking is best understood by looking at the relations between timescales, using the heuristic of
inter-scale relationships within temporal ranges. The chapter argues that simpler models of
timescales have limited explanatory use and do not naturally capture the experiential richness of
musicking. In contrast, the mapping of temporal ranges highlights the relations between many
processes that mutually enable and constrain one another across timescales, and across brains, bodies,
and environment. The map guides research into the complexity of musicking without sacri cing
disciplinary focus. It consists of three domains of organization—sensorimotor, social life, and
person/Self—interweaving ecological-enactive concepts of embodiment, self-organization,
participatory systems, attunement, normative constraints, habits, and sense-making.

Keywords: timescale, musicking, emergence, reductionism, constraint, habit, temporal range, ecological
enactive approach, self-organization, sense-making
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Introduction C10.S1

TO say that musicking occurs at many timescales is (we would hope) uncontroversial. The notion of C10.P1
timescales helps us make distinctions between phenomena as they occur within seconds, hours, years, or
centuries. Everything from transient neuronal dynamics to the ebb and ow of society and history can be
ranked according to their relative timescales. But what happens when we use timescale in a more

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
complicated way to shape how we seek scienti c explanations? For instance, at which timescale do we nd
the most fundamental explanation of musical phenomena? This question reveals con icting assumptions
about what it takes to explain music and musical behaviours. Many scholars may nd it too naive or
reductionist, while others may nd very appealing the idea of narrowing the range of scales and observable
interactions to a fundamental explanatory set. After all, the great success of physics, to which many
scientists turn for inspiration, depends on how a limited range of very fast interactions can do the heavy
lifting to explain a much wider range of phenomena spread across many timescales, from chemical bonding
to the life cycle of stars.

This chapter reworks the critique of reductionism and advances a strategy to think about complexity in C10.P2
musicking in a way that integrates both di erent research interests across the board and a global or
systemic view of musicking as a multi-temporal whole. For this strategy to work, it is crucial to see the
temporality of musicking not as a matter of di erent timescales but as a matter of more or less
interdependent timescales, and of how such interdependencies emerge.

The chapter is divided into two parts. In the rst part, ‘Mapping Timescales and Emergence’, I point to the C10.P3
p. 170 limited usefulness of describing music merely in terms of tiered timescales—say, from milliseconds to
centuries. I argue that while they may have use as descriptive models, allocating musicking processes into
tiered timescales is only provisional and not su ciently nuanced to deal with an explanatory approach to
musicking. Tiered accounts of timescales, although intuitive, are too simplistic. This limitation is in part
because key phenomena at every level of analysis spill over many timescales (Loaiza et al. 2020). The rst
part will take the reader momentarily away from common discussions in music research to rework an
argument against reductionism.

In the second part, ‘Temporal Ranges of Musicking’, I come back home to music and, with the help of an C10.P4
example from my own musical experience of becoming a batucada-style tamborim player, I elaborate a map
of what are called ‘temporal ranges’ (Ste ensen and Pedersen 2014; Loaiza et al. 2020). I propose a path
towards a naturalistic (but not reductionist) understanding of emergent organization in musicking. Distinct
domains of organization are captured through an alternative way of clustering various timescales in
temporal ranges. A map of at least three domains and their temporal ranges—sensorimotor, social life, and
personal/self—allows a characterization of key processes without severing their interdependencies. In the
second part I show how the music researcher can utilize the temporal ranges tool to draw interdependencies
across fast and slow timescales. For example, I show how to relate very fast timescales of sensorimotor
coordination (e.g. the production of the samba groove) to very slow phylogenetic timescales in a way that
allows a richer explanatory formulation.
Mapping Timescales and Emergence C10.S2

The Virtues and Vices of Tiered Timescales C10.S3

A formulation of tiered timescales is an intuitive way of describing both similar and heteronomous C10.P5
phenomena within a single frame. For example, in the list that follows we can put together a wide range of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
phenomena, including musicking, by piling up timescales progressively from ‘very slow’ (at the top of the
list) to ‘very fast’ (at the bottom of the list). This type of tiered list or mapping of timescales has descriptive
advantages and explanatory implications; however, I argue that such an approach to timescales is a
necessary but not su cient account of musicking as a complex activity:

A generic list of tiered timescales: C10.P6

a. The Universe timescale. C10.P7

b. Life on earth timescale. C10.P8

c. Human phylogenetic timescale. C10.P9

d. Historical-cultural timescale. C10.P10

e. Life cycle of an individual timescale. C10.P11

p. 171 f. Person’s social membership and social participation timescale. C10.P12

g. Music/performance/social event timescale. C10.P13

h. Micro event sequence timescale (formal substructures, gestures, rhythm/articulation, etc.). C10.P14

i. Phenomenal (self-)awareness timescale. C10.P15

j. Subpersonal/micro-behavioural timescale. C10.P16

k. Transient neurodynamic timescale. C10.P17

l. Molecular bonding timescale. C10.P18

m. Atomic timescale. C10.P19

n. Planck timescale. C10.P20

Tiered timescales are imbued in our ways of thinking about musicking through the everyday concept of C10.P21
duration. For example, the simplest formulation is to say that beats and notes are shorter than motifs,
motifs shorter than phrases, phrases shorter than pieces, and pieces shorter than concerts. Tiered
timescales are simple to use for descriptions precisely because they have duration as the point of
comparison across scales. Moreover, this kind of mapping of multiple scales in a hierarchical manner may
be used to show that many relevant events co-occur at di erent observer-dependent temporal tiers. It
shows how musicking takes time, from the duration of a single sound to historical periods. Music theory
and music history already have templates for this familiar story. First, in the way a musical piece is
analytically divided in terms of formal and rhythmic structures and subdivisions; secondly, in the way
musical performances and musical products are clustered in daily cycles, concert seasons, stylistic epochs,
musical traditions, and so on. Furthermore, tiered timescales help us show that musicking not only takes
time, it also ‘makes [our subjective] time’ (Born 2015). We can map how musicking de nes the temporality
of persons’ lives in the way individuals organize their everyday activities, life-changing events, and
autobiographies with recourse to persistent stylistic preferences and cycles of social activity (DeNora 2000).

I argue that tiered maps of timescales have descriptive power yet are poorly suited for explanatory work. A C10.P22
tiered map of timescales is an instance of a scalar hierarchy. It compares durations but does not shed light
on inter-scale relationships, especially those relationships that may have causal or explanatory signi cance
across levels.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
As a brief illustration I bring an example from my musical experience that will be expanded much further in C10.P23
the second part. My experience of learning the tamborim (a Brazilian drum) involves learning the particular
technique of the instrument ‘on the y’ while participating in an amateur group of samba batucada (Loaiza
2018). A description of becoming a tamborim player with the use of a tiered map may show processes at
many di erent timescales: a very slow process of becoming an old-timer within a community of musical
practice (over many years), a description of slow cycles of performances (following yearly seasonal events),
descriptions of preparations for particular gigs (month-length duration), event-scale descriptions of group
entrainment, joint action and attention (minutes-hours), and all the way down to the several hundred-
milliseconds scale of hands–drum coordination.

p. 172 Clearly, the number of possible tiers of description is inde nite, and surely out of reach to any single C10.P24
1
research discipline. Yet regardless of how exhaustive the description is, the inter-scale relationships
between processes remain intractable. The richness of an experience permeated by all sorts of processes
beyond the strict musical ‘here and now’ becomes, in the tiered mapping, irremediably disjointed (I return
to this shortly).

Tiered maps of timescales not only do poorly at showing inter-scale relationships, they may also re ect and C10.P25
even exacerbate pervasive disciplinary compartmentalization. Disjointed timescales are tantamount to
disjointed agendas, from empirical psychology to historical musicology.

Explanatory reductionism promises both to overcome such fragmentation and to ll the inter-scale gaps C10.P26
with the progressive discovery of how smaller and faster interactions fully explain the apparent phenomena
at scales closer to home (i.e. human life scales). I argue that this seemingly hopeful promise might in fact be
misleading. As an alternative, I point towards a richer interdisciplinary account of musicking that addresses
inter-scale relationships in a non-reductionist way. In what follows I elaborate a brief critique of
reductionism.

Rediscovering the Organizational Properties of Wholes C10.S4

Physicists have taught us that ‘the explanatory arrows always point downward’, to the very small and very C10.P27
fast (Steven Weinberg, cited in Kau man 2008: 43). Accordingly, questions about causes should always lead
us down to the fundamental spatio-temporal scales of energy-matter: from societies down to individuals,
to organs, to cells, to molecules, to atoms, and ultimately to string theory (or whichever ‘theory of
everything’ is at hand). However, while this downward explanatory strategy has had value in the history of
physics, it becomes problematic when applied to music simply because music disappears on the way down.

Let’s see where reductionism takes us. How far down is down enough to explain musicking? Based on C10.P28
current technological limitations and established disciplinary boundaries between physics and life sciences,
one may be tempted to postulate a fundamental level of musicking at the scale of neuronal activity. But if
technology can be improved and boundaries moved, why not go further down? For example, physicist Roger
Penrose (1989) proposes a further reduction of neuronal activity, seen as the stu of consciousness, to
quantum-level interactions. But if the bottom boundary is seemingly arbitrary, the question remains: how
can we know that the quantum-scale, or more controversially, the neuronal-scale are not too far down
already?

To nd out, we need to improve our questions: could one deduce the particular organization that C10.P29
corresponds to the group coordination and grooviness of a samba band from the observation of the
neuronal activity alone?

p. 173 According to complexity theorist Stuart Kau man (2008), reductionism has the insurmountable challenge C10.P30

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
of demonstrating how an unequivocal deduction can be carried out upwards from a putative causal level to a
target higher-level of phenomena, that is, to show how to go up (not just to say that ‘it is so’)—say, from
string theory to string quartets. This deduction upwards is impossible not only because its complexity is
computationally intractable but, most importantly, because the initial downward move of reduction erases
essential formal properties of the organization of energy and matter at the higher levels (organisms,
populations, etc.). Organizational (or formal) properties of wholes are not retrievable from the parts alone
and thus cannot be deduced upwards from a lower explanatory level—organization is an emergent, non-
2
reducible, phenomenon (Kau man 1993; 2008; Kirchho 2014; Thompson 2007). Turning reductionism on
its head, by rediscovering emergent levels of organization the ‘explanatory arrows’ start pointing upwards
(not only downwards) to higher-level formal constraints on wholes (Juarrero 1999).

Music seen through the lens of reductionism disappears because music’s life is dependent, not on a C10.P31
particular set of mechanisms or processes, but on the properties of the organization of many processes
across timescales. Rather than asking how far down to go (a question that nds no traction in the slippery
terrain of tiered maps of timescales and research agendas), the questions we need to ask concern the
delimitation of the relevant boundaries between domains of phenomena that manifest distinctive
organizational properties.

My proposal is to nd a principled way of making distinctions between possible organizational domains in a C10.P32
way that still speaks to di erent research interests. The challenge is complicated by the fact that musicking
(or any other human) phenomena spill over many spatio-temporal scales. For example, although we may
describe a process of becoming a skilful tamborim player in a samba band as simply happening in a slow
timescale (i.e. the move from the group’s ‘periphery’ to the ‘centre’, or from newcomer to old-timer; see
Lave and Wenger 1991), a richer explanatory account of this process will require unpacking the continuous
fast timescale re-enactment of distinctions of musical taste in the micro-feedback loops that reinforce
appropriate forms of playing, and in turn, how such loops relate to (slower changing) narratives of the self
as a member of a social group. On the one hand, as is clear by now, single-timescale explanations, for
example a social structuralist account that has no room for fast micro-sociality or individual experience,
simply do not do the job; on the other, ad hoc maps of tiered timescales may have descriptive power but no
explanatory insight.

The complexity of musicking, I suggest, requires an explanatory strategy to unpack both the distinctions C10.P33
and continuities between domains of inter-scale relationships that manifest global organizational
properties. The key idea is that several organizational domains—i.e. systems of inter-scale relationships—
overlap over many timescales, thus forming nested temporalities. This will become clear in the following
sections: my main proposal is to disentangle sensorimotor, social, and personal domains of musicking in
terms of distinct temporal ranges that represent organizational properties.
p. 174 From Tiered Timescales to Emergent Temporal Ranges C10.S5

Musicking’s organizational properties must be continuous with the organizational properties of life C10.P34
3
(Blacking 1973; Loaiza 2016; Moran 2014; Schiavio 2014; Small 1998; van der Schy 2015). I propose a
principled way of distinguishing organizational domains in musicking by recasting existing approaches to
the emergence of organization in life. Several scienti c and philosophical approaches have identi ed the
general domain of life as a strong candidate for an emergent level of natural organization (including

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
phenomenological views: Jonas and Jonas, 2001[1966]), the theory of autopoiesis (Maturana and Varela
1980), anticipatory systems (Rosen 1991), theory of complex adaptive systems (Kau man 1993),
teleodynamics (Deacon 2011), the free energy principle (Friston 2012), and organizational biology (Moreno
4
and Mossio 2015). The organization of life, in this view, is multi-scalar, comprising brain-body-
environment self-organizing systems as well as population group dynamics, and sociohistorical levels that
5
form complex wholes.

Following this line of thought, a new way of mapping timescales takes as a template the idea that the realm C10.P35
6
of life is an emergent causal domain nested within, but not reducible to, the larger realm of physics. As we
move the focus from physics to life, this nested relationship can be conceptualized as a narrowing in the
range of relevant timescales. Thus, whereas the range of physics comprises the largest timescale continuum
(between ‘Universe’ timescale and the ‘Planck’ timescale in the list above), the domain of life comprises a
narrower and more speci c timescale continuum (approximately between molecular bonding timescales
and the slow timescale of evolution).

Ecological interactivity theorists Ste ensen, Pedersen, and colleagues propose a map in which di erent C10.P36
organizational domains correspond to speci c clusters of timescales in what the authors call ‘temporal
ranges’. A temporal range is a range of timescales delimited by fast/slow timescales maxima (Loaiza et al.
2020; Ste ensen and Pedersen 2014; Uryu et al. 2013). Figure 10.1 is a formalization of the relation between
two generic temporal ranges and how they represent di erent domains with varying ranges of timescales.

The key to de ning nested temporal ranges is that the narrowing of range is simultaneously the movement C10.P37
from general to more speci c domains. This mapping procedure can be repeated to represent more speci c
7
domains depending on our focus and interests. Each added domain (towards the right hand of Figure 10.1)
is more speci c, more focused on a narrow range of timescales, and is embedded in many layers of larger,
more general domains that serve as contexts for our particular focal domain (Loaiza et al. 2020). Ste ensen
and Pedersen, taking their cue from systems theorist Stanley Salthe (2005; 1991), highlight the way a
domain or class—e.g. the domain of life—speci es or recalibrates a part of the organization of the general
physical domain, and is itself subject to further speci cation by its own subclass in a recursive way. The
result is a kind of hierarchy that is not based on scale but on a history of diachronic emergence from simpler
p. 175 organization (fewer organizational properties) to complex organization of processes (many co-
manifested or overlapping organizational properties) (Deacon 2011).
Figure 10.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Graphic formalization of temporal ranges. The vertical dimension indicates a timescale continuum from slower timescales (at C10.F1
the top of the figure) to faster timescales (at the bottom). The temporal range B (e.g. life) on the right hand in a pale grey
rectangle represents a specification or narrowing of the more general temporal range A (e.g. physics) on the le . The horizontal
dimension indicates the increase in degree of specificity. The size of each rectangle represents both the maximum range of
timescales for each temporal range and the overlapping of temporal ranges; thus, the timescales contained in B are also
contained in A. The area of B is not fully contained in A—this represents emergence of an organizational domain ʻBʼ (e.g.
emergence of life as a temporal range). Based on Ste ensen and Pedersen (2014); Uryu et al. (2013); Loaiza et al. (2020).

The narrowing of temporal ranges re ects the so-called ‘adiabatic principle’: ‘the faster that something C10.P38
happens, the less energy is transferred. Conversely, very slowly varying processes appear as a stable
8
background on the timescale of faster ones’ (Lemke 2000: 279). Following this principle, to say that life’s
domain is a subclass of the physical domain is to say that the organic domain has a relative closure that can
be rendered as being less a ected by both the fastest subatomic processes and the slowest timescales of
galactic events. Of course, living forms have necessarily evolved in conditions of, for example, gravity, and
are material things held together by subatomic forces; but the point is that the organizational domain of life
is more sensitive to fewer (not-so-fast/slow) timescales and richer in interdependent, inter-scale relationships
9
in a narrower range.

How is the recursive logic of specifying temporal ranges relevant to explaining musicking? First, a C10.P39
naturalistic approach to musicking needs to account for the continuity between biological organization and
specialized forms of social behaviour and experience. Such an approach can be formulated using the
temporal ranges logic whereby some processes within the larger temporal range of the human species (a
subclass of life), under selective pressures, have been recalibrated to form a particular genre of human
p. 176 activity. Each temporal range represents an incremental speci cation within the larger, more general
domain of life and manifests a particular observer-dependent focus. Using this non-reductionist account,
we can draw a continuity from species-generic properties to a particular focus on personal skill and
experience in musicking.
Moreover, trans-disciplinary views of musicking are hungry for the kind of step-like approach to C10.P40
organizational complexity o ered by the logic of temporal ranges and its ‘adiabatic principle’. Returning to
the musical example, as established before, a richer conceptualization of samba groove requires examining
the relationship between processes at fast and slow timescales, for example between inter-bodily real-time
entrainment and my personal membership in a community. This relationship emerges against the
background of even faster and slower processes of motor control and phylogenetic change, and
simultaneously, such a relationship itself stands as the background for richer inter-scale relationships

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
within the narrower range of my personal narratives.

The temporal ranges model o ers a simple (but not simplistic) tool to think of musicking’s rich temporal C10.P41
complexity through notions of inter-scale interdependence, emergence of organizational domains, and
nested ranges of timescales. What follows is one possible implementation of the model proposed, together
with a more detailed approach to the emergent properties at play in musicking.

Temporal Ranges of Musicking C10.S6

Basics of Mapping Musicking C10.S7

The basic map I propose, depicted in Figure 10.2, consists of three temporal ranges representing the C10.P42
domains: sensorimotor, social life, and person/Self. In what follows I ll in the details of the map by
elaborating a story that moves from more general to more speci c domains. Using an example from my
musical experience, I move from a description of a generic hand-drum activity (sensorimotor range),
through an instance of musical social activity and participation (social life range), and nally arriving at my
drumming skill in the context of my personal experience and identity as a musician in a band (person/Self
10
range).

First temporal range C10.S8

The rst temporal range (R1) (Figure 10.3) corresponds to a species-speci c organizational domain. R1 C10.P43
highlights the relation between historical processes, occurring in the very slow timescales of the species
evolution (or phylogeny), and proximate processes, occurring in the (very) fast timescales of sensorimotor
11
p. 177 (neuronal and musculoskeletal) activity.
Figure 10.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Specification map of three temporal ranges. Each rectangle with a di erent shade of grey represents a temporal range with a C10.F2
specific set of timescales (fast to slow). Timescale maxima for each range are indicated at the top and bottom of each rectangle:
e.g. the social life range is flanked by ʻmicrosocial interaction timescaleʼ and ʻsocio-historical timescaleʼ, which is read as
expressing the fastest and slowest relevant timescales of the social life temporal range. See also Uryu et al. (2013); Loaiza et al.
(2020).
Figure 10.3

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
First temporal range, R1. C10.F3

By focusing on this range we can account for the role that the basic architecture of the human body—its C10.P44
given physical biases and degrees of freedom—has in the production of musical activity—for example, the
bimanual coordination involved in hitting a drum. Such architecture is of course relative to an environment
that is also in ux and is constantly altered by human activity. In R1, the historical processes correspond to
the continuous evolutionary co-de nition and mutual shaping of body factors and environmental features
(Laland et al. 2000).

The coupling of body and environment is captured in what I call an ‘embodied relation’. In this relation the C10.P45
12
body is a kind of model of its environment (Ramstead et al. 2018). I use the idea of embodied relation as a
p. 178 heuristic for inter-scale relationships—relations between processes (historical, proximate) within a
temporal range. First, bodies must embody, in their architecture and degrees of freedom, the regularities of
their environment to engage successfully in behaviour with an environment in ux (Jordan and Day 2014).
When we look at the dispositions and physical biases of a body, we also see the re ection of the particular
conditions of its environment. For example, in the body of a tuna we see a re ection of the dynamics of open
oceanic waters in the hydrodynamic shape of the tuna’s body, the degrees of freedom of its joints, e ective
angle of vision, and so on. Secondly, it is not enough to discuss such bodily factors in terms of an animal’s
adaptations to an environment in a slow evolutionary timescale; organisms also need to ne-tune and
quickly update their embodied relation with the changing environment in faster timescales of behaviour.

In R1 the embodied relation is not based on structural similarity; rather, it is a matter of actual and possible C10.P46
dynamical correspondences, and the presence of dispositions for coupling. Human bodies embody (i.e. have
an embodied relation with) environments characterized by, among other features, the presence of detached
13
solid objects like mugs or sticks. Fingers, hands, and arms compound for a total sum of degrees of
freedom, and these possibilities of compound nger–hand–arm movements constitute an embodiment of
the myriad of possible changes of position and displacement of those detached solid objects. When nger–
hand–arm and objects interact, for example in grasping, many of such possibilities are instantiated. Under
regular conditions, objects such as mugs or sticks a ord being grabbed by hands and this tells us something
about a basic mutuality between dynamics of limbs and a world of solid objects.

Behaviour exibly interweaves in real-time matching regularities of body and environment, and the C10.P47

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
particular contingencies of an emergent body-environment dynamic. Embodied relations in real-time
interaction are always in the process of being ne-tuned while resisting dissipation or divergence. In the
case of grabbing a stick, not all movements may be possible because of the interference of other moving
objects; objects may dynamically change energy properties in unexpected ways; and human bodies across
populations present many possible variations and deviations from an idealized con guration. Moreover, in
the case of newborn babies, the coupling of limbs and objects is yet to be developed by the growing body.
This means that any su ciently successful interaction between body features and immediate environment
needs to comprise a process of compensating for divergence and the local initial conditions of suboptimal
coupling. Accordingly, I propose to characterize proximate processes in R1 as processes of reduction or
14
constraint on the divergence of the embodied relation dynamics. In this way, successfully grabbing a stick is
thus an achievement of reducing or constraining dynamical (real-time) divergence.

Constraint and the resulting reduction of divergence can be understood with dynamical concepts (Juarrero C10.P48
1999; Thompson 2007; Chemero 2009). An important assumption of dynamical approaches to behaviour is
the idea that such constrained processes are self-organized. These processes manifest the emergence of
regularities and long-range correlations without the need to account for a ‘doer’ that does the job of
p. 179 organizing and determining regularities (the doer as a putative internal or external agent with respect to
the system of interest). As a classic example of self-organizing behaviour, think of a ock of starlings and
how it can be understood without reference to a centralized command within the ock; in other words, the
ock’s behaviour is a distributed organizational property of the whole that emerges by virtue of simpler
constraining interactions between starlings. I suggest that ‘constraint’ best captures how systems do work,
progressively becoming ‘just-well-enough’ self-organized, through self-imposed limits on their degrees of
15
freedom (Deacon 2011).

In order to show the use of the heuristics presented and move on to the next temporal ranges, I provide an C10.P49
interpretation of my own experience of learning how to play the tamborim. The tamborim, as it is used in
the large samba ensembles in Rio de Janeiro, is an instrument that has the particular role of emphasizing
the samba swing. Tamborim players need to learn a basic rhythm called carreteiro which is a deceptively
‘easy’ pattern of constant sixteenths. Every third sixteenth the left hand, which holds the drum almost
upwards, turns approximately 60 degrees, so that the plastic beater (held by the right hand) hits the drum
skin in an upward movement (called virado). As expected, any novice starts with a slow tempo and then
gradually increases it. In my early practice experience my manual coordination would become unstable and
break down frequently. This occurred when reaching a certain tempo threshold (in my case roughly around
16
110 bpm). After several attempts at keeping the coordination beyond the threshold, a spontaneous change
took place. I found myself playing a pattern that seemed to ow more easily and with a highly positive
timing—call it a pleasurable grooviness. After directing my attention to the new pattern, and a few more
attempts beyond the tempo threshold, I could return to the pattern more quickly than before; it had a
‘sticky’ quality to it. I recognized the pattern as the samba batucada swing which is a micro-rhythm of four
notes per beat with an intricate arrangement of deviations from the sixteenth grid (rendering proportions
roughly closer to those of a 9/32 grid; see Naveda 2011). This was certainly a tamborim-scale ‘eureka’
moment in my learning process.

As described above, the simple event of holding and waving a drumstick shows the accomplishment of the C10.P50
body–environment coupling in a very basic sense, but it is the process of transitioning from various
unstable forms to a successfully stable pattern of drumming that calls our attention. How did this felicitous
and seemingly spontaneous transition occur? As a self-organized behaviour, the trajectory towards a less
unstable pattern corresponds to the trajectory of the system hands–stick–drum towards an attractor basin
17
in its state space. This process can be described more simply as a constrained divergence of the body–
instrument coupling whereby energetically costly or ine cient forms of interaction become discarded. The
constrained divergence manifests both an exploration and a deselection of energetically unstable forms of
interaction from a pool of all forms of theoretically possible interaction. Due to the nature of the irregular

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
energy transactions between the drum skin, the exible material of the beater, and the spring-like elasticity
of muscles and joints, we can expect that only a limited number of possible interactions are temporally
sustainable above various tempo thresholds (i.e. energy instabilities are better distributed in only a fraction
p. 180 of all possible states). As a consequence, we observe that the self-organized behaviour moves to less and
less unstable forms relative to the rate of energy input. (See Haken, Kelso, and Bunz 1985 for a paradigmatic
model of bimanual coordination.)

Proximate processes in R1 manifest constraints on the divergence of the dynamics of body and world. In C10.P51
sensorimotor interaction, the body and world tend towards optimizing their embodied relation through
self-produced constraints—we may also call this a process of sensorimotor simpli cation.

I propose that this kind of ecological and dynamical description is generalizable to many forms of C10.P52
musicking interaction. Yet the anecdote of the tamborim ‘eureka’ moment is not just about basic hands–
drum coupling (and how it re ects the species evolution in terms of selection of better- t energy
transactions); it is also about the role of a social convention regarding the appropriate swing, and my
experience of achieving the ‘right kind’ of groovy feel relative to my history as an enthusiastic listener of
Brazilian music. R1 alone o ers a general view of the link between fast timescales and very slow timescales,
yet it lacks the speci cation or depth of eld that we need. To account for the social depth on the one hand
and my personal depth on the other I introduce the two other temporal ranges.

Second Temporal Range C10.S9

In Figure 10.4, the second temporal range (R2) is represented as a smaller and more speci c range than R1: C10.P53
from microsocial-interactive timescales to sociohistorical timescales.
Figure 10.4

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Second temporal range, R2. C10.F4

R2 is a recalibration or reorganization of only a portion of the timescales of R1. R2 zooms in to the inter- C10.P54
scale relationships in the more specialized domain of social life and social interactions. Hence for the
p. 181 analysis of R2 we leave out the fast micro-changes of energy transactions occurring in sensorimotor
simpli cation and the very slow process shaping the species and its environment.

The link between sensorimotor and social life domains is a process of attunement between (at least) two C10.P55
bodies and world. Firstly, human bodies are naturally positioned in di erent degrees of disattunement with
respect to one another. Following the logic of constraint on divergence, in social interaction we may nd a
constraint on inter-body disattunement (James and Loaiza 2020). Similar to how the coupled dynamics of
hands and drums across certain tempi thresholds yields a few stable forms of drumming, the constraint on
disattunement yields less unstable inter-body interactions—i.e. it results in the selection of emergent
forms of ‘attuned sociality’ (Fuchs and De Jaegher 2009; Schutz 1976). Secondly, the constraint on
divergence in social interaction results also in a di erent attunement to the world. Think of how walking
abreast on the street while holding hands makes some movements of each person’s body unviable (changes
of direction, speed, etc.). Holding hands is a mechanical constraint on the degrees of freedom of the inter-
body coupling (resulting in bodies located in close range) and the degrees of freedom of each body relative
to the street.

Attunement, moreover, is also constrained by what seem to be external or trans-situational factors (Cowley C10.P56
and Ste ensen 2015; Steiner and Stewart 2009). When we walk on the street, we are also constrained by
tra c lights, yet clearly there is nothing intrinsic to the wavelength of the colour red that determines an
immediate reduction of one’s viable movements and thus preventing one from crossing the street. In such
situations one’s attention is not only directed towards pragmatic (mechanical) aspects of the world; it is
also oriented towards elements of normative interest.

Social normative constraints are by de nition trans-situational, and thus they are di cult to capture in the C10.P57
narrow timescales of social interaction. How can we talk about social normativity in a more naturalistic
way? A rst hint towards tackling the puzzle of normative constraints is to see how the suggested bottom-
up transition from R1 to R2 is not only a conceptual construct but also an observed phenomenon in real-life
ontogeny. The developmental transition of newborns into intersubjective attention and further into
intersubjective situations including objects of interest (see Trevarthen 1979 regarding primary and
secondary intersubjectivity) is also the transition into social regimes of orientation of attention via the
interactive play of pointing to objects and events of interest (see Tomasello 2008). The e cacy of the red
light on my body is rst (developmentally) discovered through my attunement with the body and gaze of
the other person as a way of being attentionally oriented through pointing-like acts. Crucially, ‘pointing’
can also be formulated for sounds and movement. I elaborate brie y this dynamical and phenomenological

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
discussion with the use of the tamborim example.

Social Interaction and Normative Constraints C10.S10

R2 highlights the link between proximate processes characterized by the normative orientation of attention C10.P58
in social interaction and the historical processes that shape larger groups of people and their normative
p. 182 biases in slower timescales (e.g. institutions, markets, communities of practice). At this point I propose a
crucial conceptual move: in R2 the heuristic of an embodied relation serves to describe the relation between
an interpersonal participatory system and a world of social practices and relationships. That is, a
participatory system embodies its sociocultural world. Embodiment, at this point, is not predicated on the
individual organism (or agent) but on the participatory relationship and dynamics comprising two (or
more) individuals in any speci c interactive situation. The couple holding hands and walking, a tango dance
couple, or a jazz ensemble are examples of participatory and situated systems (De Jaegher and Di Paolo
2007; Kyselo and Tschacher 2015; Walton et al. 2015). A participatory system is, I propose, the appropriate
level of analysis for proximate (relatively fast timescales) processes in R2.

Historical processes, on the other hand, correspond to the slow shaping of the multiplicity of sociocultural C10.P59
regularities that are evident in the distribution of tasks and roles, the partition of private and public spaces,
the interconnection of artefacts and their uses, amongst other macro-scale organizational forms (and as far
as the global economy). These forms are presented as crystallized structures yet belong nonetheless to a
continuous self-organizing process of history and economic or productive relations (Kau man 2008).

Focusing on the proximate processes, the proposal consists of describing the constraint on the divergence C10.P60
of a participatory embodied relation. We may think of two kinds of constraint processes closely intertwined
in R2. Firstly, as aforementioned social interaction can be seen as a constraint on inter-body disattunement.
This constraint results in the emergence of local and temporal intrinsic patterns; in a way, it can be said to
constitute an autonomous self-sustaining system (Auvray and Rohde 2012; De Jaegher and Di Paolo 2007;
Froese et al. 2014). For example, forms of interpersonal musical entrainment manifest such emergent
intrinsic patterns (Clayton 2012). Secondly, social interactions also manifest dynamical forms of orientation
of attention in a pragmatic ecology (Gibson 1969; Ingold 2000; van den Herik 2018). Persons in interaction
shape each other’s attentional trajectories through bodily changes and utterances, and in the process,
attention can be directed towards salient objects, events, and other occurrences in the interactive context.

18
To clarify this, let me bring back the tamborim example. Suppose that I ask an experienced surdo player to C10.P61
help me with my tamborim practice session. The surdo player would play a simple rhythm of quarter notes
in cut time. Accordingly, after initially establishing a slow practice tempo, we begin to accelerate together
without any external help. Yet sometimes, because my tamborim playing is not yet stable, the coordination
tends to break, sometimes it breaks completely, and sometimes the coordination is recovered by returning
uninterruptedly to slower tempi. Our movements and feel for the interaction are tense. Later on, in the
middle of our coordination, a sudden change in my tamborim playing is then matched by a distinctive
gesture of my fellow surdo player (e.g. a head movement and a grin). After this, the coordination seems to
stabilize in a pleasurable samba batucada groove (as already described). Our movements feel more uid and
p. 183 we can enjoy staying in a relatively stable upbeat tempo for longer. We then try to repeat the process and
nd out that every iteration is slightly easier—we achieve the groove faster than before in each try.

From this example we can distinguish several dynamical and phenomenological aspects. First, in order to C10.P62
observe the e ect of a constraint on disattunement, I propose that the tempo acceleration needs to be
understood as a single joint or co-de ned action (Di Paolo et al. 2018). That is, an action that is best de ned
with two levels in mind: a global control or organizational level of a situation, and the level of coupled (but

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
not fused) co-operant interactors whose trajectories momentarily become heavily constrained by the global
organization. Thus, rather than observing an instance of two juxtaposed systems—a leader and a follower—
externally engaged in a sequence of turn-taking actions and reactions, what we observe in joint (co-
de ned) action is the emergent trajectory of a single, global dynamical system of temporally coupled
interactors (Cummins 2013). In this way, the successful acceleration itself (for the time it is maintained) is a
partial achievement of the constraint on dis-attunement of the whole system (surdo player, myself, and
drums), not of each separate individual. This type of constraint process de nes the participatory system’s
boundaries in fast timescales. The participatory system engenders an intrinsic pattern manifested in the
characteristic spontaneity that drives the interaction (De Jaegher and Di Paolo 2007; Walton et al. 2015).
Secondly, as already pointed out, the interaction e ects a selective process by means of the orientation of
attention towards very particular forms available through joint action. Bodily changes in my fellow surdo
player may mark transitions proper to our trajectories of attention. Such bodily changes constitute actions
of pointing to ‘phenomenal objects of interest’ (Hutchins 1995). In this case, the pointing is normative; it
corresponds to a positive normative interpretation. Through their bodily changes (e.g. face expression) the
surdo player orients the attention towards an event that has some valence, and through their pointing I see
that I nally ‘got the groove’ in my tamborim. In this way, I nd myself immersed in a normative play in
which my attention uctuates and is pragmatically oriented by the attention of the other, not only in their
gaze but also with the conduct of their bodies, gestures, facial expressions, and utterances.

Taken together, the constraint on dis-attunement and the constraint realized by the orientation of C10.P63
attention enable and reinforce each another in ways that vary according to the common microhistory of
persons in interaction. The result of orientation of attention is a kind of second-order reduction of the pool
of possible correlations between the intrinsic states of the participatory system and features/events of the
world. The pattern of orientation of attention, with its normative interpretations, compounds with the
constraint on disattunement (reduces the amount of viable joint actions); simultaneously, the successful
achievement of joint action makes possible the incorporation of a normative constraint. In other words, we
can nd the appropriate groove in our interaction thanks to the exercise of a normative interpretation, and
reciprocally the normative orientation of attention gains traction by virtue of the constraint on inter-body
disattunement.

As implied above, normative interpretations make reference to a trans-situational framework, but such a C10.P64
p. 184 framework is not causally e cient without a myriad social interactions occurring all the time and
everywhere. Hence, although we can, for example, construct an analytical or rhetoric gloss of the correct
samba swing on paper (in terms of music theory or ethnomusicology), it is through the incorporation of the
normative constraint into the live dynamics of social interaction that constraints are held in place.
Normative constraints are thus constraints that both exist ‘on their own’ in a slower timescale—as we
capture them in our descriptions of social practice—and also need to be (re-)enacted in the faster
interactions in everyday life for their existence. Thus, from the perspective of the incorporation of
normative constraints into fast timescales, we can make sense of the idea that an interpersonal
19
participatory system embodies (has an embodied relation to) its sociocultural world. Crucially, how this
can be achieved consistently across time, place, and di erent people is a matter of the redundancy of
constraints in the physical medium and the tendency to habituation in social life (James and Loaiza 2020).
On the one hand, the historical processes in R2 shape the physical world by rendering it informationally C10.P65
20
redundant. Tra c lights, for example, can carry a very limited amount of (theoretical) information by the
fact of always displaying the same number of colours in alternation. Consistent correlations between
changes of states in the world (e.g. changes of light wavelength) and the changes of orientation of attention
in social interaction are in part maintained simply by keeping available information (e.g. possible colour
states) relatively stable across generations. In the case of a tra c light it is part of an engineered niche, but
in many other cases it is the e ect of some form of social inertia (self-organization plays a role in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
describing macro-social behaviour). For our musicological case, we can always count on the material
regularities of instrumentation, notation, distribution of private and public spaces, as well as the cycles of
musicking, (e.g. festival season) that keep the average environmental information entropy low across time.
The study of such material/niche regularities corresponds to an extended theory of a ordances (Rietveld
and Kiverstein 2014; in music, see Clarke 2005 and Krueger 2014). On the other hand, the possibility of
existing communal practices of attention and normative interpretation depends on the general disposition
of individuals and larger groups towards incorporating habits. I propose that habit can be seen as a pivotal
concept linking R2 and the third temporal range (R3).

Transition from R2 to R3: Habits C10.S11

As aforementioned, the transition from R1 to R2 corresponds to attunement in social acts. Now the C10.P66
transition from R2 to R3 involves the concept of habit. These transitional notions o er interdisciplinary
perspectives, for example, the notion of habit may be articulated both at the social and personal levels of
analysis, thus amounting to, for example, the synthesis of sociological and idiographic insights. Figure 10.5
represents these transitions within the proposed map.

Returning to the tamborim example, it is clear that at some point in my history of tamborim-learning, the C10.P67
p. 185 socially embedded act of pointing to the groove was a necessary event. But this act is of course not
subsequently required in every single instance of my tamborim playing; in fact, I can play a groovy pattern
alone in the privacy of my home and be conscious of my own playing in the absence of the other’s
attentional orientation and participatory dynamics. In short, I can describe this phenomenon by saying that
I have acquired dispositions and habits of playing, listening, and (self-)awareness (Becker 2004; Turino
2008).
Figure 10.5

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Transitions from R1 to R2 and from R2 to R3. C10.F5

Habits can be said to be persistent patterns of action that criss-cross the boundaries between individual, C10.P68
21
environment, and collective. The concept points to the emergence of self-sustaining patterns that
integrate nested levels of constrained dynamics: from various cooperating lower level sensorimotor forms
of organization, e.g. entrainment (Barandiaran and Di Paolo 2014; Egbert and Barandiaran 2014), through
person-level and interpersonal forms of identity (James and Loaiza 2020), to the global or sociocultural
level of tendencies and distinctions of taste (Bourdieu 1977). Accordingly, habit emphasizes the circular
causality of many processes at di erent timescales, and thus provides a tangible approach to the more
abstract concept of dynamical emergence of organization.

Habit conceptually bridges the gaps between the immediacy of sensorimotor engagements, the mediation C10.P69
by means of conventions and stabilities in the social world, and the persistence of a person’s dispositions or
tendencies to action and interpretation. Thus, to be able to describe a habit amounts to capturing patterns of
organization in both personal and collective levels of analysis. In this way, distinctions of taste, the uses of
language, the life cycles of musical styles, the fetishization of artefacts and technology, the performance of
gender, amongst many examples, manifest constellations of habits that coordinate collective and personal
processes.

In order to make the transition from R2 to R3, the focus now is on how only a reduced set of habits can C10.P70
cluster together and persist over time as a coherent bundle which constitutes an individuated and Self-
ascribing person. The transition represents the idea that persons may be seen as reorganizations of some
habits from a pool of public (socially available) habits. Such habits are not simply piled up randomly; they
p. 186 enable and e ect selective pressures onto one another. An account consisting of cooperating and
mutually enabling/constraining habits compressed into the narrower timescales of an individual
organism’s history may bring light to the process of individuation of a musicking Self.

Third Temporal Range C10.S12

Finally, the third temporal range (R3) (Figure 10.6) highlights the continuity between a person’s life cycle C10.P71
and the person’s sense-making.
Figure 10.6

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Third temporal range, R3. C10.F6

Sense-making, as a proximate process ‘here and now’, is an embodied and situated regulation of patterns of C10.P72
activity, including a ective-emotive, cognitive, and bodily patterns (Colombetti 2013; Ste ensen 2015;
Thompson 2007). Crucially, sense-making is a function, not of the implementation of context-independent
and internal cognitive mechanisms, but of the personal history of habitual patterns of action, a ect,
attention, and interpretation. Tying this view of sense-making with the embodied relation heuristics:
sense-making embodies an individual’s life history and habitual patterns. In this way, ‘here and now’
activity manifests a constraint on the divergence of a person’s embodied (a ective and cognitive) history.

In R3, historical processes correspond to ontogeny understood not so much as a development towards a C10.P73
nal stage of physiological maturity but as an ongoing process of growth and pattern stabilization (Smith
and Thelen 2003). Ontogeny extends along the full life cycle of the individual and comprises physiological as
well as social, experiential and a ective trajectories (Lewis 2000). Just as physiological growth is the
ongoing self-construction—through cellular di erentiation, reproduction, and repair—of a coherent whole
precariously distinguished from the physico-chemical environment, social and experiential growth is the
p. 187 ongoing incorporation and preservation of a coherent set of habits that allows the individual to be both
distinct from the biosocial environment and adaptable to it (James and Loaiza 2020).

Ontogeny is the ratchet-like process that secures and preserves constraints that come to be embodied as C10.P74
personalized habits (i.e. habits of speech, emotional pro les, etc.); it manifests through forms of socio-
experiential identities (Froese and Di Paolo 2011). Social life, through cycles of social interaction, propels
this process. In the course of a person’s life the individual is pulled into participatory systems, and the
increasing participation into larger systems of actions in socioculture. Think, for example, of how adults
when interacting with preverbal children ‘pull them in’ into participatory systems by addressing children
with repetitive vocal gestures (e.g. wordings) before the child can even match those same speech-producing
movements (van den Herik 2021). The subsequent habituation of the child to the challenges of such
situations corresponds to the harnessing of a lower level of constraint on interactional divergence through a
higher-level tendency towards increasing participation in a micro-community (i.e. a family). The emerging
speech habits of the child not only allow simple communication but also manifest the consolidation of an
identity (as a talking member of the family, for example). Moreover, they enable further exposure to even
more complex forms of participation. By securing and preserving patterns of action at relatively lower
levels, it is possible to exploit new dimensions of participation o ered in social life at relatively higher
levels. Ultimately, the person is not only ‘pulled in’ but also comes to ‘push on’ to maximize access to
resources only available through increasing participation (Loaiza 2019). Ontogeny tightly interweaves
increasing participation in social life and the construction of identity (Lave and Wenger 1991; Vygotsky
2012[1962]). For verbal children and adults this also comprises forms of explicit identity through linguistic

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
habits of self-designation, self-interpretation, and styles of declarative memory.

My participation in a community of samba music practice presents a miniature version of the ontogenetic C10.P75
22
process. As I grow within the community of practice, I acquire both an implicit and an explicit form of
identity as a tamborim player (Lewis and Ferrari 2001). The correct production of grooviness in the
tamborim corresponds to the (re-)enactment of sedimented (implicit) and interdependent habits of
listening, orientation of attention, and sensorimotor simpli cation. The resonance of this grooviness in the
other musicians’ bodies facilitates joint access to more specialized sociocultural resources such as samba
festivals and important gigs: we become a samba band known for having a tight groove. At the personal
level this participatory move is re ected in forms of explicit identity. I can now designate myself as a skilled
tamborim player with recourse to narratives borrowed from the folk repertoire of stories of musicianship.
Crucially, the fact of this self-designation does not consist so much of a simple statement about the
existence of skills and habits of tamborim playing (and the extent that this claim can be true) but instead
manifests a habitual use of wider-ranging narratives about my Self and the reproduction of a personal
autobiography. I keep track and amplify an existing personal narrative whilst constraining the divergence of
storytelling with respect to my musical skills—for example, being a South American musician, exposed
p. 188 since childhood to the social life and the feel of Afro-Latin musics, I feel closer to home in my own hand
movements and body swing while playing samba. In this way, I’m not only reporting a state of a airs, I’m
also (and most importantly) reconstructing my Self with the use of my own repertoire of narratives.

Thus, the embodied relation in R3 has an explicit version as a model or representation of the Self (Metzinger C10.P76
2015). Thanks to our constant immersion in intersubjective practices in social life, we are able to maintain,
update, and transform that model or representation of the Self through borrowing from a shared repertoire
of narratives (see Sutton et al. 2010; Hutto 2008). Those narratives tell us how to (re)construct a state of
a airs and also help us keep track of our personal autobiographies. In sum, we place ourselves in the
narratives available in the social life we share with others; in doing so, we re-present our Selves to ourselves
and others (Mead 2009[1932]; Di Paolo et al. 2018; Thompson 2014). Thus, the focus of the description of
the explicit model is on the enaction of patterns of habitual utterances that reconstruct in their occurrence a
particular model or representation of the Self.

The particularity of proximate sense-making processes in R3 is that persons enact a model of their Selves C10.P77
through the use of habitual practices, including narrative practices, available in musicking. Given the
tendency of habits to self-sustain, the expectation is that those habits can be conserved in the future.
Sense-making thus conveys a temporal extension towards a projected future—it brings a temporal
experience of anticipation. In this sense, the model is related to expectations over the course of actions in a
particular situation. Sense-making, through the use of the explicit model, thus has epistemic consequences:
it consists of the anticipation of a future state given a prior expectation of conservation of identity and the
necessary actions (including wordings/utterances) that constrain or reduce uncertainty under contingent
conditions. Against the background of social life, in which inde nite degrees of freedom expand the horizon
of uncertainty, the enactment of a Self stands as an island of reduced uncertainty through the intertwining
of sensorimotor, inter-bodily, attentional, and narrative constraints.
Concluding Remarks C10.S13

To say that music, from sounds and movement to traditions and heritage, constitutes a complex whole is C10.P78
(we would hope) uncontroversial (Merriam 1964). An intuitive way to approach complexity is to expand the
temporal horizon of our observations and distinguish between tiers in a vast array of timescales, while
capturing the dynamics of networks and processes operating in each tier. Although such temporal

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
expansion points in the right direction and enriches our multidisciplinary discourses, it falls short of
solving a crucial problem that frameworks based on mapping tiered timescales must face—how to explain
and do justice to the experiential richness of musicking without eliminating or severing the inter-scale
interdependencies at play. Prevailing reductionism o ers a costly solution: it denies altogether the
explanatory role of inter-scale relationships while seeking to atten the analysis of complex phenomena to
p. 189 a narrow range of scales, usually to the fastest and smallest scales possible. But seen through the lens of
reductionism, music disappears; it is not explained, but explained away.

I have argued in favour of an explanatory target ultimately consisting of the organizational properties of C10.P79
musicking, which are themselves a version of those properties we nd in life. Like life, music is dependent,
not on a particular set of mechanisms or processes, but on the properties of the organization of many
processes that mutually enable and constrain one another across timescales. As I have shown, the seemingly
simple activity of getting the samba groove right in my own instrumental practice turns out to involve a
coherently integrated meshwork of inter-scale relationships—local and non-local factors across bodies and
environment, as well as proximate and historical processes, individual and collective, spilling over many
timescales. The task of the researcher is to spell out such inter-scale relationships as a means to
approximate their implicated organizational properties.

I have proposed a map as a guide to researching the complexity of musicking without sacri cing C10.P80
disciplinary focus and perspective. It consists of at least three temporal ranges that represent domains of
organization in musicking—sensorimotor, social life, and personal/Self. The map makes use of tiered
timescales in a restricted sense to highlight the emergence, from more general to more speci c, of inter-
scale interdependencies within ranges of timescales. It is an e ort to generate a non-reductionist yet
naturalistic view of the emergence of organizational properties by showing both continuities and
distinctions between domains that are best described with the logic of nested temporal ranges.

I have suggested recruiting various concepts existing in the ecological-enactive cognition literature that C10.P81
capture di erent stages, sizes, and facets of relationships: embodied relation, participatory system, habit,
and narrative model of Self. A process of constraint or reduction of divergence operates across these forms
of relationship, securing their convergence to regular dynamical paths. Beyond these suggestions, the
mapping proposed lays out the ground and leaves open the discussion of alternative con gurations of inter-
23
scale relationships, and especially the discussion of open-endedness and creativity in musicking.

Notes
1. In the practice, the number of tiers under consideration would be determined by the availability of measuring devices and C10.N1
their resolution, since each process/timescale would be described as data sets given data points with particular intervals
in their capture.

2. These authors, amongst others, have defended emergentism as diachronic, dynamical, and process-based, which is C10.N2
di erent to classic emergentism. Deacon (2011) also defends a view of diachronic emergence of form-as-cause, but o ers
a constructive critique of the use of the notion of organization (as it is indeed too observer-dependent).

3. Musicking is a phenomenon of life, and thus its study needs to be informed by systemic approaches to the continuity of C10.N3
lifeʼs organizational phenomena, while overcoming orthodox cognitive computational/representational, and so-called
p. 190 neo-Darwinian views. I bring together aspects of ecological and enactive approaches to mind (Chemero 2009; He
2001; Thompson 2007; Varela et al. 1991), an ethological approach to the free energy principle (Ramstead et al. 2018), and
process-based philosophy of biology (Nicholson and Dupre 2018). In ecological and enactive approaches, cognition,
a ectivity, and social life are distributed processes not reducible to internal mechanisms of the individual organism.

4. The idea of a distinctive organization in life goes back to Kant (Critique of Judgement, pt 2), who observed the C10.N4
organizational closure of whole organisms in contrast to machines.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
5. This contrasts with its opposite extreme idea in reductionism that the organization of life is nothing but the existence of C10.N5
order in a genome.

6. To be clear, no laws of physics are harmed in the making of this chapter. Otherwise put, at no point is there any suggestion C10.N6
that life is outside physics or that life violates any of the well-established laws of physics (e.g. the second law of
thermodynamics). Nor is there any appeal to mysterious vital forces.

7. Formalizing the representation in Figure 10.1, the relation of A and B takes the form {A{B}}, where B specifies A. To define C10.N7
even narrower temporal ranges in a recursive fashion we may obtain {A{B{C{…}}}}, where e.g. {physical domain {organic
domain {human species {ontogeny {specific focus of interest}}}}} (Salthe 2005). Salthe (2005) makes an important
distinction between Scalar Hierarchies (I call them tiered maps) and Specification Hierarchies.

8. In Lemkeʼs words: ʻAdiabatic basically means “it doesnʼt get through,” referring to energy, fields, or informationʼ (2000: C10.N8
279).

9. Catastrophic events of course do alter life dramatically in multiple timescales, from cosmic rays to quantum disruptions. C10.N9
Yet such external perturbations are events, not processes within lifeʼs interdependent cycles.

10. It is possible to fill the map in the opposite direction by considering how specific domains alter their relative contextual, C10.N10
more general domains. That is, it is possible to go from right to le and not only le to right, as I do herea er.

11. I use the distinction between ʻproximateʼ and ʻhistoricalʼ processes. These concepts reinterpret two notions widely used in C10.N11
discussions about causality, namely, proximate and ultimate causes (see Ramstead et al. 2018). First introduced to
evolutionary biology by Mayr (1961) and elaborated by ethologist Tinbergen (1963: see ʻfour questionsʼ), proximate and
ultimate causes roughly refer to ʻhowʼ and ʻwhyʼ questions respectively. For example, investigating the attraction of some
bacteria towards light, we may explicate the mechanisms linking photoreceptors and the activation of flagella (i.e. a
proximate causal story, or ʻhowʼ), as well as elaborate on the possible evolutionary history of selection that rendered
bacterial phototaxis adaptive (i.e. an ultimate causal story, or ʻwhyʼ). Proximate views tend to focus on explaining
mechanisms instantiated in the individual organism, whereas ultimate views usually correspond to statistical
explanations of population-level phenomena (Ariew 2003). In my interpretation, the relation of ʻhowʼ and ʻwhy as well as
the emphasis on statistical path selection are preserved; what changes is the way we think of each temporal range as
having relative proximate and historical processes.

12. The authors propose an information-theoretic concept of generative model: ʻan organism does not just encode a model of C10.N12
the world, it is a model of the world—a physical transcription of causal regularities in its eco-niche that has been sculpted
by reciprocal interactions between self-organization and selection over timeʼ (Ramstead et al. 2018: 5).

13. This idea of embodied relation, seen from ecological psychology, can be interpreted in terms of a ʻlandscape of C10.N13
a ordancesʼ whereby a ordances are always relational in a form of life (Rietveld and Kiverstein 2014).

p. 191 14. The use of the word ʻreductionʼ here must not be confused with previous references to reductionism as a methodological C10.N14
and philosophical approach. ʻReductionʼ here is used in a technical sense to describe tendencies of coupled systems to
limit the variety of their states. Bruineberg and Rietveld (2014) first introduced the term in a similar way. For them, the
process consists of the optimization of grip on a field of a ordances. See note 13.

15. See a more technical elaboration of constraint in Loaiza et al. (2020). The notion of reduction or constraint on divergence C10.N15
may be seen as bringing together aspects of qualitative definitions found in the Dynamical Systems Theory literature
(DST) (Kelso 1995; Chemero 2009) and concepts widely used in the music literature such as resonance, coordination,
synchronization or entrainment (Clayton 2012).
16. Of course, this could be addressed more empirically. I show this as an anecdote to illustrate the discussion. C10.N16

17. State space and attractor are DST concepts. A state space is defined by all the possible states of the system. The attractor is C10.N17
a point in the state space that describes a tendency.

18. The surdo is a Brazilian low-pitched drum. C10.N18

19. The use of embodiment is not of course intended to capture all aspects of the interpersonal participatory system. Such C10.N19
systems may well modify in unpredictable ways the larger social medium. An elaboration of the creative potential of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
participatory systems is outside the scope of the present discussion.

20. Here I use information only in Shannonʼs sense. The ecological theory of a ordances (Chemero 2009) expands on the use C10.N20
of the notion of information.

21. ʻHabitʼ has a long history in philosophy and psychology: e.g. William James described habit as the unit of mental life. The C10.N21
term became disused as a consequence of the emergence of strong cognitivism in the 1950s (especially with Chomsky).
My use also takes what Pierre Bourdieu called ʻhabitusʼ: ʻsystems of durable, transposable dispositions, structured
structures predisposed to function as structuring structures […] being all this, collectively orchestrated without being the
product of the orchestrating action of a conductorʼ (Bourdieu 1977: 72).

22. Following the logic of temporal ranges, it could be rightly argued that a particular example of musical practice constitutes C10.N22
a subset of temporal relationships, i.e. a further, narrower temporal range of its own. My elaboration of the temporal
ranges tool leaves open precisely this possibility. See Loaiza et al. (2020).

23. Two hints for an approach to natural open-endedness: one lies in the concept of constraint itself, which should not be C10.N23
understood strictly in a deterministic sense but should include spontaneous mutations. Each instance of constraint on
divergence implies a series of transitions from unstable activity to simplified and stable interactions—starting with a pool
of possible yet unstable configurations and evolving towards a more constrained (less varied) domain of viable
alternatives. However, the exact target configuration of a constrained process at each level cannot be determined a priori
because a stable interaction that is aptly selected can be instantiated by many possible alternative configurations. Any
transition thus requires a spontaneous exploration of many possible configurations, and some of these may constitute
completely new stable forms. In short, constraint processes resulting in emergent organization, at whichever level, have a
degree of open-endedness. The other hint is the way of reading the map itself, which I have only explored from general to
specific. Having in mind the mutability of constraints, we could think of cascades of mutations propagating in both
directions, not only from the general to the specific.
p. 192
References C10.S14

Ariew, A. (2003). Ernst Mayrʼs ʻultimate/proximateʼ distinction reconsidered and reconstructed. Biology and Philosophy 18: 553– C10.P82
565.
Google Scholar WorldCat

Auvray, M., and Rohde, M. (2012). Perceptual crossing: The simplest online paradigm. Frontiers in Human Neuroscience 6: 181. C10.P83

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
https://doi.org/10.3389/fnhum.2012.00181
Google Scholar WorldCat

Barandiaran, X. E., and Di Paolo, E. (2014). A genealogical map of the concept of habit. Frontiers in Human Neuroscience 8: 522. C10.P84
https://doi.org/10.3389/fnhum.2014.00522
Google Scholar WorldCat

Becker, J. O. (2004). Deep listeners: Music, emotion, and trancing. Indiana University Press. C10.P85
Google Scholar Google Preview WorldCat COPAC

Blacking, J. (1973). How musical is man? Faber & Faber. C10.P86


Google Scholar Google Preview WorldCat COPAC

Born, G. (2015). Making time: Temporality, history, and the cultural object. New Literary History 46(3): 361–386. C10.P87
https://doi.org/10.1353/nlh.2015.0025
Google Scholar WorldCat

Bourdieu, P. (1977). Outline of a theory of practice (trans. R. Nice). Cambridge University Press. C10.P88
Google Scholar Google Preview WorldCat COPAC

Bruineberg, J., and Rietveld, E. (2014). Self-organization, free energy minimization, and optimal grip on a field of a ordances. C10.P89
Frontiers in Human Neuroscience 8: 599. https://doi.org/10.3389/fnhum.2014.00599
Google Scholar WorldCat

Chemero, A. (2009). Radical embodied cognitive science. MIT Press. C10.P90


Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (2005). Ways of listening: An ecological approach to the perception of musical meaning. Oxford University Press. C10.P91
Google Scholar Google Preview WorldCat COPAC

Clayton, M. (2012). What is entrainment? Definition and applications in musical research. Empirical Musicology Review 7(1-2): 49– C10.P92
56.
Google Scholar WorldCat

Colombetti, G. (2013). The feeling body: A ective science meets the enactive mind. MIT Press. C10.P93
Google Scholar Google Preview WorldCat COPAC

Cowley, S. J., and Ste ensen, S. V. (2015). Coordination in language temporality and time-ranging. Interaction Studies 16(3): 474– C10.P94
494.
Google Scholar WorldCat

Cummins, F. (2013). Joint speech: The missing link between speech and music? Percepta 1(1): 17–32. C10.P95
Google Scholar WorldCat

De Jaegher, H., and Di Paolo, E. (2007). Participatory sense-making. Phenomenology and the Cognitive Sciences 6(4): 485–507. C10.P96
https://doi.org/10.1007/s11097-007-9076-9
Google Scholar WorldCat
Deacon, T. W. (2011). Incomplete nature: How mind emerged from matter. Norton. C10.P97
Google Scholar Google Preview WorldCat COPAC

DeNora, T. (2000). Music in everyday life. Cambridge University Press. C10.P98


Google Scholar Google Preview WorldCat COPAC

Di Paolo, E., Cu ari, E. C., and De Jaegher, H. (2018). Linguistic bodies: The continuity between life and language. MIT Press. C10.P99
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Egbert, M., and Barandiaran, X. E. (2014). Modelling habits as self-sustaining patterns of sensorimotor behavior. Frontiers in C10.P100
Human Neuroscience 8: 590. https://doi.org/10.3389/fnhum.2014.00590
Google Scholar WorldCat

Friston, K. (2012). A free energy principle for biological systems. Entropy 14: 2100–2121. https://doi:10.3390/e14112100 C10.P101
Google Scholar WorldCat

Froese, T., and Di Paolo, E. (2011). The enactive approach: Theoretical sketches from cell to society. Pragmatics and Cognition C10.P102
19(1): 1–36. https://doi:10.1075/pc.19.1.01fr
Google Scholar WorldCat

Froese, T., Iizuka, H., and Ikegami, T. (2014). Embodied social interaction constitutes social cognition in pairs of humans: A C10.P103
minimalist virtual reality experiment. Scientific Reports 4: 3672. https://doi.org/10.1038/srep03672
Google Scholar WorldCat

Fuchs, T., and De Jaegher, H. (2009). Enactive intersubjectivity: Participatory sense-making and mutual incorporation. C10.P104
Phenomenology and the Cognitive Sciences 8(4): 465–486. https://doi.org/10.1007/s11097-009-9136-4
Google Scholar WorldCat

Gibson, E. J. (1969). Principles of perceptual learning and development. Appleton-Century-Cro s. C10.P105


Google Scholar Google Preview WorldCat COPAC

p. 193 Haken, H., Kelso, J. A. S., and Bunz, H. (1985). A theoretical model of phase transitions in human hand movements. Biological C10.P106
Cybernetics 51(5): 347–356.
Google Scholar WorldCat

He , H. (2001). Ecological psychology in context: James Gibson, Roger Barker, and the legacy of William Jamesʼs radical C10.P107
empiricism. Psychology Press.
Google Scholar Google Preview WorldCat COPAC

Hutchins, E. (1995). Cognition in the wild. MIT press. C10.P108


Google Scholar Google Preview WorldCat COPAC

Hutto, D. D. (2008). Folk psychological narratives: The sociocultural basis of understanding reasons. MIT Press. C10.P109
Google Scholar Google Preview WorldCat COPAC

Ingold, T. (2000). The perception of the environment: essays on livelihood, dwelling and skill. Psychology Press. C10.P110
Google Scholar Google Preview WorldCat COPAC

James, M., and Loaiza, J. M. (2020). Co-enhabiting interpersonal inter-identities in recurring social interaction. Frontiers in C10.P111
Theoretical and Philosophical Psychology 11: 577. https://doi.org/10.3389/fpsyg.2020.00577
Google Scholar WorldCat

Jonas, H., and Jonas, E. (2001). The phenomenon of life: Toward a philosophical biology. Northwestern University Press. C10.P112
Google Scholar Google Preview WorldCat COPAC
Jordan, J. S., and Day, B. (2014). Wild systems theory as a 21st century coherence framework for cognitive science. MIND Group. C10.P113
Google Scholar Google Preview WorldCat COPAC

Juarrero, A. (1999). Dynamics in action. MIT Press. C10.P114


Google Scholar Google Preview WorldCat COPAC

Kau man, S. A. (1993). Origins of order: Self-organization and selection in evolution. Oxford University Press. C10.P115
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Kau man, S. A. (2008). Reinventing the sacred: A new view of science, reason, and religion. Basic Books. C10.P116
Google Scholar Google Preview WorldCat COPAC

Kelso, J. S. (1995). Dynamic patterns: The self-organization of brain and behavior. MIT Press. C10.P117
Google Scholar Google Preview WorldCat COPAC

Kirchho , M. (2014). In search of ontological emergence: diachronic, but non-supervenient. Axiomathes 24(1): 89–116. C10.P118
Google Scholar WorldCat

Krueger, J. (2014). A ordances and the musically extended mind. Frontiers in Psychology, 4: 1003. C10.P119
https://doi.org/10.3389/fpsyg.2013.01003
Google Scholar WorldCat

Kyselo, M., and Tschacher, W. (2015). An enactive and dynamical systems theory account of dyadic relationships. Frontiers in C10.P120
Psychology 4: 452. https://doi.org/10.3389/fpsyg.2014.00452
Google Scholar WorldCat

Laland, K. N., Odling-Smee, J., and Feldman, M. W. (2000). Niche construction, biological evolution, and cultural change. C10.P121
Behavioral and Brain Sciences 23(1): 131–146.
Google Scholar WorldCat

Lave, J., and Wenger, E. (1991). Situated learning: Legitimate peripheral participation. Cambridge University Press. C10.P122
Google Scholar Google Preview WorldCat COPAC

Lemke, J. L. (2000). Across the scales of time: Artifacts, activities, and meanings in ecosocial systems. Mind, Culture, and Activity C10.P123
7(4): 273–290. https://doi.org/10.1207/S15327884MCA0704_03
Google Scholar WorldCat

Lewis, M. D. (2000). The promise of dynamic systems approaches for an integrated account of human development. Child C10.P124
Development 71(1): 36–43.
Google Scholar WorldCat

Lewis, M. D., and Ferrari, M. (2001). Cognitive-emotional self-organization in personality development and personal identity. In C10.P125
H. A. Bosma and E. S. Kunnen (eds), Identity and emotion: Development through self-organization, 177–198. Cambridge University
Press.
Google Scholar Google Preview WorldCat COPAC

Loaiza, J. M. (2016). Musicking, embodiment and participatory enaction of music: Outline and key points. Connection Science C10.P126
28(4): 1–13. https://doi.org/10.1080/09540091.2016.1236366
Google Scholar WorldCat

Loaiza, J. M. (2018). Humaning through music: An enactive approach to the continuity of mind and life with implications for C10.P127
musicking. Doctoral dissertation, Queenʼs University Belfast.
Google Scholar Google Preview WorldCat COPAC

Loaiza, J. M. (2019). From enactive concern to care in social life: towards an enactive anthropology of caring. Adaptive Behavior C10.P128
27(1): 17–30.
Google Scholar WorldCat

p. 194 Loaiza, J. M., Trasmundi, S. B., and Ste ensen, S. V. (2020). Multiscalar temporality in human behaviour: A case study of C10.P129
constraint interdependence in psychotherapy. Frontiers in Psychology 11: 1685. https://doi.org/10.3389/fpsyg.2020.01685
Google Scholar WorldCat

Maturana, H., and Varela, F. J. (eds) (1980). Autopoiesis and cognition: The realization of the living. Reidel. C10.P130

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Mayr, E. 1961. Cause and e ect in biology. Science 134(3489): 1501–1506. C10.P131
Google Scholar WorldCat

Mead, G. H. (2009 [1932]). Mind, self, and society: From the standpoint of a social behaviorist. University of Chicago Press. C10.P132
Google Scholar Google Preview WorldCat COPAC

Metzinger, T. K. (2015). M-autonomy. Journal of Consciousness Studies 22(11/12): 270–302. C10.P133


Google Scholar WorldCat

Merriam, A. P. (1964). The anthropology of music. Northwestern University Press. C10.P134


Google Scholar Google Preview WorldCat COPAC

Moran, N. (2014). Social implications arise in embodied music cognition research which can counter musicological C10.P135
individualism. Frontiers in Psychology 5: 676. https://doi.org/10.3389/fpsyg.2014.00676
Google Scholar WorldCat

Moreno, A., and Mossio, M. (2015). Biological autonomy: A philosophical and theoretical enquiry. Springer Science. C10.P136
Google Scholar Google Preview WorldCat COPAC

Naveda, L. (2011). Gesture in samba: A cross-modal analysis of dance and music from the Afro-Brazilian culture. Doctoral C10.P137
dissertation, Ghent University.
Google Scholar Google Preview WorldCat COPAC

Nicholson, D. J., and Dupré, J. (eds) (2018). Everything flows: Towards a processual philosophy of biology. Oxford University Press. C10.P138
Google Scholar Google Preview WorldCat COPAC

Penrose, R. (1989). Emperorʼs new mind: Concerning computers, minds and the laws of physics. Oxford University Press. C10.P139
Google Scholar Google Preview WorldCat COPAC

Ramstead, M. J. D., Badcock, P. B., and Friston, K. J. (2018). Answering Schroedingerʼs question: A free-energy formulation. C10.P140
Physics of Life Reviews 24: 1–16.
Google Scholar WorldCat

Rietveld, E., and Kiverstein, J. (2014). A rich landscape of a ordances. Ecological Psychology 26(4): 325–352. C10.P141
Google Scholar WorldCat

Rosen, R. (1991). Life itself: A comprehensive inquiry into the nature, origin, and fabrication of life. Columbia University Press. C10.P142
Google Scholar Google Preview WorldCat COPAC

Salthe, S. N. (1991). Two forms of hierarchy theory in western discourses. International Journal of General Systems 18(3): 251– C10.P143
264.
Google Scholar WorldCat

Salthe, S. N. (2005). Two frameworks for complexity generation in biological systems. In C. Gershenson and T. Lenaerts (eds), C10.P144
Evolution of complexity: ALifeX proceedings, 99–104. Indiana University Press.
Schiavio, A. (2014). Action, enaction, inter(en)action. Empirical Musicology Review 9(3): 254–262. C10.P145
Google Scholar WorldCat

Schutz, A. (1976). Making music together: a study in social relationship. In Alfred Schutz: Collected papers, vol. 2: Studies in social C10.P146
theory, ed. A. Brodersen, 159–178. Martinus Nijho .
Google Scholar Google Preview WorldCat COPAC

Small, C. (1998). Musicking: The meanings of performing and listening. Wesleyan University Press of New England. C10.P147

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Smith, L. B., and Thelen, E. (2003). Development as a dynamic system. Trends in Cognitive Sciences 7(8): 343–348. C10.P148
Google Scholar WorldCat

Ste ensen, S. V. (2015). Distributed language and dialogism: notes on non-locality, sense-making and interactivity. Language C10.P149
Sciences 50: 105–119.
Google Scholar WorldCat

Ste ensen, S. V., and Pedersen, S. B. (2014). Temporal dynamics in human interaction. Cybernetics and Human Knowing 21(1/2): C10.P150
80–97.
Google Scholar WorldCat

Steiner, P., and Stewart, J. R. (2009). From autonomy to heteronomy (and back): The enaction of social life. Phenomenology and C10.P151
the Cognitive Sciences 8(4): 527–550.
Google Scholar WorldCat

p. 195 Sutton, J., Harris, C., Keil, P., and Barnier, A. (2010). The psychology of memory, extended cognition, and socially distributed C10.P152
remembering. Phenomenology and the Cognitive Sciences 9(4): 521–560. https://doi.org/10.1007/s11097-010-9182-y
Google Scholar WorldCat

Thompson, E. (2007). Mind in life: Biology, phenomenology, and the sciences of mind. Harvard University Press. C10.P153
Google Scholar Google Preview WorldCat COPAC

Thompson, E. (2014). Waking, dreaming, being: Self and consciousness in neuroscience, meditation, and philosophy. Columbia C10.P154
University Press.
Google Scholar Google Preview WorldCat COPAC

Tinbergen, N. (1963). On aims and methods of ethology. Zeitschri für Tierpsychologie 20: 410–433. [This journal was renamed C10.P155
Ethology from 1986.]
Google Scholar WorldCat

Tomasello, M. (2008). Origins of human communication. MIT Press. C10.P156


Google Scholar Google Preview WorldCat COPAC

Trevarthen, C. (1979). Communication and cooperation in early infancy: A description of primary intersubjectivity. Before Speech C10.P157
1: 530–571.
Google Scholar WorldCat

Turino, T. (2008). Music as social life: The politics of participation. University of Chicago Press. C10.P158
Google Scholar Google Preview WorldCat COPAC

Uryu, M., Ste ensen, S. V., and Kramsch, C. (2013). The ecology of intercultural interaction: Timescales, temporal ranges and C10.P159
identity dynamics. Language Sciences 41: 41–59.
Google Scholar WorldCat

van den Herik, J. C. (2018). Attentional actions: An ecological-enactive account of utterances of concrete words. Psychology of C10.P160
Language and Communication 22(1): 90–123.
Google Scholar WorldCat

van den Herik, J. C. (2021). Rules as resources: an ecological-enactive perspective on linguistic normativity. Phenomenology and C10.P161
the Cognitive Sciences 20(1): 93–116. https://doi.org/10.1007/s11097-020-09676-0
WorldCat

van der Schy , D. (2015). Music as a manifestation of life: Exploring enactivism and the Eastern perspective for music education. C10.P162

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469567 by National Science & Technology Library user on 26 May 2023
Frontiers in Psychology 6: 345. https://doi:10.3389/fpsyg.2015.00345
Google Scholar WorldCat

Varela, F. J., Thompson, E., and Rosch, E. (1991). The embodied mind: Cognitive science and human experience. MIT Press. C10.P163
Google Scholar Google Preview WorldCat COPAC

Vygotsky, L. S. (2012 [1962]). Thought and language (ed. A. Kozulin; trans. E. Hanfmann and G. Vakar). MIT Press. C10.P164
Google Scholar Google Preview WorldCat COPAC

Walton, A. E., Richardson, M. J., Langland-Hassan, P., and Chemero, A. (2015). Improvisation and the self-organization of multiple C10.P165
p. 196 musical bodies. Frontiers in Psychology 6: 313. https://doi.org/10.3389/fpsyg.2015.00313
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
CHAPTER

11 Understanding Musical Instants  C11

Rolf Inge Godøy

https://doi.org/10.1093/oxfordhb/9780190947279.013.9 Pages 197–C11.P135


Published: 08 December 2021

Abstract
We may typically experience music as continuous streams of sound and associated body motion, yet we
may also perceive music as sequences of more discontinuous events, or as strings of chunks with
multimodal sensations of sound and body motion, chunks that can be called ‘sound-motion objects’.
The focus in this chapter is on how such sound-motion objects emerge at intermittent points in time
called ‘musical instants’, and how musical instants are necessary in order to perceive salient features
in music such as of timbre, pitch, texture, contour, and overall stylistic and a ective features. The
emergence of musical instants is also understood as based on the combined constraints of musical
instruments, sound-producing body motion, and music perception, also suggesting that
understanding musical instants may have practical applications in making music.

Keywords: continuity, discontinuity, intermittency, body motion, sound-motion objects, constraints


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

Introduction C11.S1

THE focus in this chapter is on our subjective experiences of brief events in music, of fragments of sound and C11.P1
associated body motion, typically in the 0.3–3 seconds duration range. Although we regard music as
unfolding in time, the claim here is that fragments of music at this timescale may be perceived and
conceived holistically and ‘in a now’, to borrow an expression from Edmund Husserl (1991), hence, as what
may be called ‘musical instants’.

Musical instants can include any kind of sound fragments, instrumental, vocal, electronic, or C11.P2
environmental, and be singular (e.g. a trumpet tone, a single guitar pluck, a hammer hitting a metal rod) or
composite (e.g. contain any kind of rhythmic, textural, melodic, harmonic, and timbral pattern), provided
that such fragments are short and are perceived holistically as somehow coherent entities (e.g. as a fanfare
motif, a rapid harp arpeggio, a waltz pattern, or a wah-wah sound of the opening and closing a trombone
mute). Musical instants are crucial for our experience of music, because in spite of their rather short
duration, they may be su cient for perceiving not only basic musical features such as timbre, dynamics,
and pitch, but often also more high-level salient musical features such as style, aesthetics, sense of motion,
and a ect, due to the cumulative and ‘instantaneous’ presence in our minds of sequentially unfolding sound
and associated body motion events.

Yet the idea of musical instants is challenging because events that have temporal extension in our C11.P3

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
environment, i.e. sequentially unfolding sound and associated body motion, are transformed in our minds
into something perceived as instantaneous, into seemingly more solid sound and motion images. How
events that unfold in time can be recalled as instantaneous overview images is a conundrum that has
p. 198 received much attention in phenomenological philosophy and gestalt theory. The remarkable insights
from some late nineteenth- and early twentieth-century thinkers in these domains on the coexistence of
continuity and discontinuity in our experience of the world can now be extended with more recent ndings
in cognitive psychology, in particular in human movement science, with insights about intermittency in
motor control and e ort. Intermittency is here understood as the constraint-based disposition of human
behaviour for discontinuous, point-by-point schemes of motion control and perception. And this is then
the main idea of the present chapter: to try to understand the subjective experiences of musical instants as
grounded in the predispositions of our organism for intermittent cognition.

To arrive at such an understanding of musical instants, we need to start out with some considerations of C11.P4
what seems to be a spontaneous and instantaneous perception of signi cations in listening (covered in the
next section, ‘Listening’). Considering elements of listening will be necessary when we go on to an overview
of the timescales involved in the generation and perception of di erent musical features, e.g. duration
thresholds for perceiving salient features such as pitch, timbre, texture, style, and sense of motion
(‘Timescales’). The topic of timescales in music is in turn closely linked with various constraints, both of
our bodies and our minds, in the generation of music (‘Constraints’). Combined elements of timescales and
production constraints will then in turn be necessary for understanding intermittency in music
(‘Intermittency’), as well as for understanding how perceptual and motor-related features in music may be
thought of as shapes, as indeed instantaneous overview images of that which unfolds in time such as a
melodic contour or a timbral transition (‘Shapes’). This all converges in what may be called ‘sound-motion
objects’, meaning multimodal fragments of sound and body motion that may be experienced holistically as
musical instants (‘Sound-Motion Objects’). Finally, it will be concluded that an enhanced understanding of
musical instants may be useful in performance, improvisation, and composition, but that there are many
important challenges ahead in securing a more solid basis for understanding the workings of musical
instants (‘Conclusions and Prospects’).

Listening C11.S2

It is useful to start by looking at some elements in the process of listening that seem to contribute to the C11.P5
sensation of musical instants. As suggested by the ecological approach to auditory event perception (Gaver
1993), we may tend to spontaneously assign labels to whatever it is that we are hearing in our environment,
typically concerning causality and signi cance, e.g., ‘bell’, ‘dog’, ‘car’, ‘approaching train’. Such auditory
event perception seems not to require much re ection as to what we are hearing, nor will there usually be
much focus on the sound features as such. Rather, the sound signal, or the ‘carrier’ of the event
information, tends to ‘disappear’ the moment the meaning of the event has been extracted and processed
further in our minds in interaction with pre-existing schemas and categories. By such fast and holistic
p. 199 perception of auditory events, it seems that we have ‘auditory instants’ in everyday listening, and that
this seems to be a spontaneous and e ortless capability for perceiving and thinking, similar to the fast
1
System 1 of Daniel Kahneman (2011).
But, as documented in auditory scene analysis research (e.g. Bregman 1990), event perception may require C11.P6
the sound to have a gestalt-like coherent ordering of low-level spectral features in the signal, as well as
providing access to cues at higher levels of organization (often also to cues from other sense modalities), in
order to work well. Hence, sound events are ontologically composite in the sense that they have multiple
components, ranging from low-level physical signal features to various higher-level signi cations. This
ontological complexity also raises issues of top-down schematic processing vs bottom-up emergence in
auditory perception, and the general view now seems to be that of a dynamical interaction between these

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
two streams (Bizley and Cohen 2013; Gri ths and Warren 2004).

Given the top-down in uence in listening, it means that our attitude, or what has come to be called our C11.P7
‘intentionality’ in listening, may be decisive for how we perceive sound. An interesting model for this was
proposed by Pierre Schae er in connection with the development of his theory of so-called ‘sound objects’
of musique concrète (1966). Sound objects in Schae er’s theory are fragments of sound from a variety of
sources, e.g. musical instruments, human voice, sound synthesis, or the environment, typically in the 0.3–3
second duration range. The use of sound objects was initially a pragmatic tool for composition in musique
concrète, consisting of mixing the playback of looped fragments on phonograph disks. However, Schae er
understood that after countless repetitions, the composers’ attention tended to be diverted from the more
immediate and/or everyday signi cations of the sound objects (cf. the notion of an ecological approach
already mentioned) to the more sonic-perceptual features of such looped sound fragments. This shift of
attention, within Schae er’s model, came to be called ‘reduced listening’, and signi ed an intentional
2
shifting of focus that is also an important feature of phenomenological philosophy (Husserl 1982).
Crucially, this method of intentional listening enhanced the awareness of salient sonic features such as the
overall dynamic and pitch-related shapes, what came to be called the ‘typology of the sonic objects’, and the
multidimensional web of internal timbral-textural features, what came to be called the ‘morphology of the
sonic objects’ (see ‘Shapes’).

The typology concerns the energy envelope of the sound, and can be seen to correlate closely with images of C11.P8
body motion, e.g. the envelope of a percussive sound corresponding with an impulsive body motion (more
on this in ‘Constraints’). For this reason, it seems intuitive to extend the idea of sonic objects to also include
corresponding body motion, hence, the extension of Schae er’s concepts to ‘gestural sonorous objects’
(Godøy 2006). This extension was based on the so-called ‘motor theory’ perspective on perception
(Galantucci et al. 2006), speci cally on an adaptation of this theory that I have called ‘motormimetic
cognition’, signifying that images of sound-producing body motion are closely linked with images of the
sonic objects (Godøy 2001; 2003).

This means that images of sound-producing body motion are often re-enacted in our minds when we listen C11.P9
p. 200 to, or merely imagine, music. This is based on the fact that musical sound is (traditionally) included in a
body motion trajectory, such as hitting, stroking, bowing, blowing, or singing. In this perspective, a sound
event does not only include the audible components, but also the motion trajectory images of the sound-
producing e ector, e.g. the hand from the initial position to the impact with the drum, and the continuation
after the impact with a rebound, and then a return to the starting point, or on to a new drum-hitting motion
sequence.

In addition to mental images of such sound-producing body motion, motormimetic cognition in listening C11.P10
may also include so-called ‘sound-accompanying body motion’, such as dancing, walking, or gesticulating
to the music. Music seems to trigger a number of body motion responses, and hence the idea of multiple
gestural a ordances of musical sound (Godøy 2010a). In many cases there will be a similarity of shape
between the dynamic envelope of the sound and the evoked body motion shape in the mind of the listener,
e.g. a sforzato chord in an orchestra evoking the image of a rapid downward hand motion. Furthermore, such
schemas may become generic and applicable across similar instances in listening. Thus, perceived patterns
of sound-accompanying body motion may contribute to our perceptions of musical sound patterns in
listening (Chemin et al. 2014).

Importantly, this converges in understanding sound and body motion chunks as related in view of gestalt C11.P11
principles (cf. the basic similarities between gestalts in perception and motor control: Klapp and Jagacinski
2011). Such chunks of sound and body motion, what I shall call ‘sound-motion objects’, have features in
parallel that are fused into coherent entities, entities that may be subjectively perceived as musical instants

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
(see ‘Sound-Motion Objects’).

Timescales C11.S3

We may experience music as a continuous stream, as an uninterrupted ow of sound and body motion C11.P12
sensations. Yet in the course of the unfolding music, we may shift our attention to certain just passed events
in the music (e.g. to a high-pitched tone on a trumpet, to a drum ll, to a screaming guitar solo), as
suggested by both the event-based and sound object-based accounts of listening already mentioned. But
knowing that musical features are based on temporal unfolding, we may also realize that our perceptions in
listening are based on cumulative and atemporal assessments of that which is inherently time-dependent.

A possible solution to this conundrum of temporal continuity and discontinuity in experiences of music C11.P13
could then be to di erentiate various timescales involved in music. This means trying to classify salient
perceptual features in music in view of their typical durations. The main questions will then be: what are
typical minimum durations necessary for perceiving salient features such as the loudness, timbre, or pitch
of any musical sound? Or for perceiving the textural, rhythmical, or melodic patterns, as well as the
p. 201 a ective and motion-related features of any musical fragment? Clearly, this will create a rather lengthy
list of feature-related timescales, extending from that of single impulses and vibrations in the sub-ms
range to large-scale formal features in the range of several minutes and even hours. But to have a more
concise overview of timescales, I have, in previous research on music-related sound and body motion,
found it practical to distinguish three main timescales categories in view of their salient perceptual features
(Godøy 2010b):

• Micro, with continuous features such as pitch, dynamics, and quasi-stationary timbre. These features C11.P14
may typically need very short durations in order to be clearly perceived, i.e. sometimes (dependent on
content and context) down to the 20 ms range and even lower (Suied et al. 2014).

• Meso, meaning fragments typically in the 0.3–3 seconds range as mentioned earlier, usually with C11.P15
clearly perceivable dynamic and pitch-related envelopes, as well as rhythmical, textural, melodic, and
harmonic patterns. In some cases, fragments with durations in the 300 ms range may be coarsely
identi ed (Gjerdingen and Perrott 2008), again suggesting that factors of content and context play an
important role here. Signi cantly, the lower duration range here in listening (≈300 ms) has also been
suggested as the typical duration for intermittent motor control (more on this later).

• Macro, meaning several meso-level chunks in concatenation, constituting periods, sections, and C11.P16
whole works. This is not a well-researched area, but it seems that perception of such large-scale forms
may allow more variation (i.e. alternative orderings of the elements) than the two other timescales
(Eitan and Granot 2008).

These timescales are concurrent, i.e. they coexist in the music and are interdependent in the sense that there C11.P17
will often be contextual e ects both upwards and downwards, e.g. that micro timescale features are found at
the meso timescale, and that the meso timescale context modi es the perceptual framework for the micro
timescale. As listeners, we may also intentionally zoom in and out of details, shifting our attention to
di erent timescales. This is what Schae er (1966) denoted with the twin terms ‘context’ and ‘contexture’
in his sound object research, meaning that a given feature has both a context (e.g. a single tone in the middle
of a phrase with several tones) and a contexture (e.g. the internal unfolding of the tone from the attack point
to the last part of the decay), and that the interplay of these timescales is a crucial factor of musical
experience.

In musical imagery (Godøy 2001), we seem to have the possibility, at will, to zoom between di erent levels C11.P18

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
of resolution, and also shift between slow motion and instantaneous reenactments, enabling what Paul
Hindemith seemed to have in mind with the following: ‘If we cannot, in the ash of a single moment, see a
composition in its absolute entirety, with every pertinent detail in its proper place, we are not genuine
creators’ (Hindemith 2000: 61).

This means that although the above-mentioned features are time-dependent, we may also think of these C11.P19
p. 202 features qualitatively as atemporal, such as in assigning pitch and interval labels, rhythmical duration
labels, and formal labels. As pointed out by Xenakis (1992) with his concepts of en-temps (in time) and hors-
temps (outside time), we may freely move between depicting temporal unfolding and more atemporal
features when we think about music. This was in particular evident in Xenakis’s own musical work
demonstrating a close a nity between musical sound and more atemporal architectural shapes, such as in
his orchestral work Metastasis and the architectural design of the Philips Pavilion at the World’s Fair in
Brussels in 1958, where whole passages of orchestral sound are indeed re ected as instants in the graphics
of the score, as what we will think of as the shape element of musical instants.

Interestingly, similar ideas of stepping out of time were previously discussed by Husserl and some of his C11.P20
contemporaries, leading up to Husserl’s notion of ‘now-points’ (Godøy 2008; 2010c; 2011; Husserl 1991;
Schneider and Godøy 2001). For Husserl, these now-points were a necessity in order to extract any meaning
at all, because without the stepping out of the continuous ow of time, there would only be an
undi erentiated stream of sensations. The phenomenological view of discontinuity seems to agree with
several other publications suggesting the need for perceptual coherence at the ‘moment’ timescale (e.g.
Michon 1978 or Pöppel 1997).

In summary, we have concurrent timescales in music perception and imagery, i.e. micro timescale features C11.P21
embedded in meso timescale features, and these in turn embedded in macro timescale contexts, and they
may all be part of our subjective experiences of musical instants. However, the meso timescale will be the
most important here due to various constraints: physical constraints of both sound and instruments, and
constraints of sound-producing body motion and of perception, requiring us now to look at how these
constraints contribute to experiences of musical instants.

Constraints C11.S4

The repertoire of musical sound made by traditional (i.e. pre-electronic) means is indeed very great, yet C11.P22
there are also limitations on what various instruments can output, as well as limitations on what human
performers can do. These limitations are the basis for several constraints that are interesting here because
they contribute to the formation of musical instants in the form of what could be called ‘musical quanta’,
meaning sound objects with salient shapes in their actual physical unfolding, such as the sound of a plucked
guitar string, of a cymbal struck with a stick, of a rapid trombone glissando (Godøy 2013).

The point of mentioning constraints here is not in any way to reduce the value of music, but rather to C11.P23
recognize that music, as an intrinsic human activity, is in fact piggybacking on certain constraints of
instruments, as well as of our bodies and minds. It seems that we are so familiar with how constraints shape
p. 203 music that we actually expect to nd these constraints when we experience music. First, there are
constraints of the instruments and sound:

• Instruments have a limited repertoire of possible sounds because of their construction, e.g. a plucked C11.P24
instrument can produce di erent kinds of pluck sounds, but not a bowed or blown sound (unless some
additional tools are used).

• Sound is typically constrained by the combination of the instrument’s physics and the human C11.P25

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
excitatory and modulatory motion on the instrument in question (or on the human vocal apparatus),
e.g. usually limited in duration, event density, loudness, etc. because of human physical limitations.

• Propagation of sound, including space, rooms, medium, etc. also constrain sound (e.g. set limits to how C11.P26
far and for how long a sound may resonate), may tend to colour the sound, and tend to smear
sequences of distinct sounds.

As for sound-producing body motion, there are some signi cant constraints: C11.P27

• Human body motion takes time, i.e. there is no instantaneous displacement of the sound-producing C11.P28
e ectors ( ngers, hands, arms, feet, lips, tongue), hence there will always be motion trajectories
between di erent postures in sound production, resulting in so-called ‘coarticulation’ (more on this
below).

• Musicians need to optimize motion. To avoid strain injury and to enable sustained performance, C11.P29
musicians will try to minimize the e ort in sound-producing motion by reducing the amount and
duration of muscle contraction, exploiting rebounds and conserving momentum.

• Motor control requirements, in particular of highly demanding musical passages, will typically need C11.P30
preprogramming and hierarchical control, as there is just not enough time to make decisions
continuously (Grafton and Hamilton 2007). In particular, the so-called ‘psychological refractory
period’, meaning the shutting o of new motor control messages within a certain timespan, seems to
necessitate gestalt-like holistic conception and control in skilled body motion (Klapp and Jagacinski
2011).

• Slowness of our motor control system. Now we see research suggesting that there is intermittency in C11.P31
motor control, meaning an uneven distribution of control, in a point-by-point kind of impulse-driven
control, or by what has been called ‘serial ballistic control’ (Loram et al. 2014), because our motor
control system is too slow for continuous feedback control. This also means that in the phases between
these intermittent control points, the e ectors are left to carry out the motion without interference
until the next intermittent control point, as is typical of so-called ballistic motion, e.g. hitting, kicking,
scraping, rapid stroking, etc.

• Motor control by key-postures, meaning optimizing planning and control of motion by preprogrammed C11.P32
shape and position of e ectors ( ngers, hands, arms, mouth, etc.) at salient moments in time,
p. 204 typically at downbeats and other accented points, and having continuous trajectories in the form of
‘pre xes’ and ‘su xes’ between these key-postures (Godøy 2008). Derived from the use of so-called
‘key-frames’ and ‘interframes’ in animation (Rosenbaum et al. 2007; Rosenbaum 2017), key-postures
may serve as orientation landmarks in time and space, and are particularly interesting in
understanding the emergence of musical instants.

Furthermore, there are constraints of sound-producing body motion in the interaction with the C11.P33
instruments, resulting in biomechanically fairly distinct categories as well as correspondingly distinct
sound categories:
• Impulsive: typically plucking, hitting, rapidly stroking, i.e. motion with burst of e ort followed by C11.P34
relaxation as in ballistic body motion, and resulting in a percussive or plucked sound.

• Sustained: blowing, bowing, slowly stroking, hence, motion with more continuous e ort, resulting in C11.P35
more protracted sound.

• Iterative: shaking, trembling, rapid back-and-forth stroking or rolling, resulting in various kinds of C11.P36
tremolo and trills.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
These are basic categories, and are usually mutually exclusive because of corresponding biomechanical C11.P37
constraints (e.g. impulsive motion is ballistic and discontinuous, whereas sustained motion is continuous).
But there are also other salient emergent features due to body motion constraints:

• Phase-transitions, meaning that there may be a change of motion category by changes in the duration C11.P38
and rate of motion elements. For instance, if a singular impulsive motion is progressively lengthened,
it may change into what is sensed as a continuous motion; conversely, if a continuous motion is
progressively shortened, it will turn into an impulsive motion; and if we have a series of well-spaced
impulsive motions which are becoming progressively more densely spaced, these may turn into an
iterative motion (cf. an illustration of this in Figure 11.1). Such category changes due to incremental
changes in one or more of its parameters is referred to as ‘phase-transition’ (Haken et al. 1985), and is
one of the most important principles of chunking motion, hence also for the sensation of musical
instants.

• Coarticulation, meaning a similar fusion of small-scale motion elements into larger-scale motion and C11.P39
sound chunks, emerging here, however, from motion between a series of key-postures as
schematically illustrated in Figure 11.2. The principle of coarticulation is well known in linguistics
(Hardcastle and Hewlett 1999) and various elds of human movement science (Rosenbaum 2010), and
stems from the fact that immediate displacement of e ectors is not possible (i.e. that all human body
motion takes time), and that there will be transition motion between key-postures. For instance:
p. 205 ngers cannot move instantaneously from one piano key to another and will often also require that
the whole hand, and even arm and shoulder, move in preparation for the next key depression. Or: the
shape of the mouth cannot change instantaneously from that of one vowel to a di erent vowel but will
require a more gradual transition from one shape to another. This means that there will be a contextual
smearing of motion in coarticulation, both backwards, called ‘carry-over e ects’, and forwards,
‘anticipatory e ects’. In addition, there will be so-called ‘spatial coarticulation’ in the sense that an
e ector may recruit motion of several other e ectors, e.g. nger motion in piano performance
recruiting hand, arm, shoulder, and even whole-body motion. Thus, coarticulation means activation
spreading both backwards and forwards in time as well as to other parts of the e ector set. It seems
that coarticulation may contribute to create coherence both in the sound-producing motion and in the
resultant sound in music (see Godøy 2014).
Figure 11.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
Schematic illustration of phase-transition by decrease of inter-onset distance. Three initially distinct impulsive events in (a), C11.F1
when moved closer together as in (b), will result in the new back-and-forth motion trajectory in (c), hence making a qualitative
change from impulsive to iterative motion as in (d).

Common to phase-transition and coarticulation is the creation of coherence at the chunk level by fusion of C11.P40
otherwise separate elements into gestalt-like units. This may then in turn be seen as the source of emergent
sensations of musical instants.
p. 206 Figure 11.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
Schematic illustration of one single key-posture with its prefix and su ix, hence with motion returning to the starting point in (a), C11.F2
and of four intermittent key-postures each with likewise prefixes and su ixes in (b). What happens when these four key-postures
are moved closer in (c) is that a new and undulating motion trajectory emerges due to coarticulation as in (d).

Intermittency C11.S5

From these elements of listening, timescales, and constraints in music, we arrive at the key concept of C11.P41
intermittency in music-related body motion. The concept of intermittency has a background in
phenomenology, in particular in Husserl’s above-mentioned idea of discontinuity in time perception
(Husserl 1991); but a remarkably similar line of thinking has been emerging in recent years from the domain
of human motor control.

For a start, it has been claimed that human motor control is anticipatory, that it works by pre- C11.P42
p. 207 programming, or in a feedforward manner (called ‘open loop’ control), because in skilled and fast body
motion there is not enough time for continuous corrective feedback (for so-called ‘closed loop’ control).
The relative sluggishness of our organism’s motor control is thus seen as the direct cause of intermittent
motor control, and as resulting in our organism instead working by a series of control impulses, or in a
point-by-point manner by the serial ballistic control mentioned earlier, and with pre-programmed body
motion chunks (Karniel 2013; Loram et al. 2014; Sakaguchi et al. 2015). Although the neurophysiological
workings of intermittency remain to be better understood, relevant research seems to o er clear
behavioural arguments in favor of intermittency in body motion generation. It may be hypothesized that
intermittency in control also entails intermittency of e ort, because control cannot be separated from
energy-requiring motion, i.e. there can be no ‘pure’ control, analogous to ‘power steering’ in machine
control systems, in human body motion. There is evidence from electromyographic (EMG) recordings of
muscle activity that e ort is not equally distributed throughout body motion chunks (Aoki et al. 1989). And
in line with the mentioned motormimetic approach to music perception, it is hypothesized that this
intermittency extends to music perception with projections of schemas from sound-producing body motion
onto whatever sound we are hearing, contributing to the formation of sound-motion objects (see ‘Sound-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
Motion Objects’).

To document the existence of these preprogrammed chunks and their triggering in music-related body C11.P43
motion, it is now feasible to study the motion trajectories of these chunks using motion capture
technologies, and to document the trajectories of e ector displacement (of ngers, hands, arms, etc.) in
relation to the key-postures on an instrument, i.e. points in the motion of particular signi cance for motor
control (cf. the presentation of key-postures in ‘Constraints’). It is also possible to calculate the derivatives
of this e ector motion, primarily velocity and acceleration (Hogan and Sternad 2007). Typically, points of
velocity reversals, e.g. at the striking point in a percussive motion where the motion direction suddenly
changes, may indicate intermittent key-posture points (Godøy 2013).

In addition, it is possible to study the e ort in generating these body motion trajectories using EMG C11.P44
recordings of muscle activation to learn more about the temporal distribution of such e ort. Using EMG is
particularly interesting for detecting the non-active segments in time, so-called ‘pre-motion silent
periods’ (Aoki et al. 1989), so as to get a clearer impression of discontinuities of e ort.

Another interesting feature of intermittency in body motion is that it may be ‘hidden’ in the sense that C11.P45
dense sequences of intermittently activated motion chunks may appear as continuous streams of body
motion. This has been referred to in Gawthorp et al. (2011) as the ‘masquerading property of intermittent
control’, and ts well with the above-mentioned idea of concurrent timescales, i.e. that shorter duration
chunks are nested within longer duration concatenated sequences of chunks. This does not change the
status of intermittency here, but it sheds light on the conundrum of the continuous and discontinuous
elements in musical experience: multiple concurrent timescale layers may be at work, but (as already
argued) the meso timescale is privileged due to the multiple convergent constraints of our bodies and
minds.

p. 208
Shapes C11.S6

A further common feature of much recent research on musical sound and on human body motion is that of C11.P46
representing various features as shapes: concerning sound, we have shapes of the unfolding sound in the
so-called ‘time domain’, at di erent levels of resolution ranging from the waveform to more superordinate
envelopes, and in the so-called ‘frequency domain’ of spectral content, we may have shape images of both
stationary and more transient spectral components; and in human movement science, we typically have
shape representations of postures and of motion trajectories, but also of derivatives showing crucial
elements such as the velocity, and the velocity reversals, already mentioned. The common feature of these
shape representations is that they capture that which unfolds in time in more atemporal, instantaneous
images. Said di erently: Shapes are inherently instantaneous in the sense of being geometric entities that we
subjectively perceive, or have in our eld of vision, ‘in a now’ and ‘all at once’.

Although we might think of shapes as primarily in the visual domain, they have extensive applications in a C11.P47
general and amodal sense. An important contribution of research on shapes in human cognition has
emerged from so-called ‘morphodynamical theory’ (Godøy 1997; Petitot 1990; Thom 1983), with the basic
principle that shapes are inherently holistic and can give us a better understanding of various phenomena in
our world than more abstract, symbol-based representations. To envisage a complex phenomenon as a
shape is not only a matter of scienti c method but may also be an intrinsic feature of perception and
cognition in general, as is evident from the large number of shape metaphors used in most areas of human
experience.

And because of their generic nature, shapes may mediate between the modalities, e.g. shapes of sound- C11.P48
producing motion may be correlated with shapes of pitch contours. How our minds make such connections

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
between di erent modalities remains to be further explored; but we can observe that the use of verbal shape
metaphors and graphical shape representations, ranging from score notation to more signal-based shape
representations of features, is so widespread in music-related contexts that there can be little doubt about
the pragmatic utility of shape images in music (Godøy 2017). Readily available technologies enable the
exploration of shape correlations of most sound and body motion features, for instance attack transient
shapes, formantic shapes, articulation shapes, and various pitch, dynamic, and timbral nuance shapes, with
body motion-related shapes such as postures, motion trajectories, motion hierarchies, or rhythmical
patterns. We can thus think of two main classes of shapes in our context, those of music-related body
motion, and those of music-related sound features. For body motion shapes, we typically have the
following:

• Key-posture e ector shapes, e.g. of ngers/hands for playing chords, of mouth for formants in vocal C11.P49
sound.

p. 209 • E ector motion shapes, e.g. hand moving across keyboard with ngers playing an arpeggio, mouth C11.P50
changing shape from vowel to vowel.

• Derivatives of motion, i.e. shapes of velocity and acceleration. C11.P51

These shapes may be anticipatory or retrospective, may include both quasi-stationary key-posture and C11.P52
body motion trajectory shapes, but are typically ‘all at once’, and hence, instantaneous. As for music-
related sound features, we often nd the following shapes:

• Dynamic contours: crescendo, decrescendo, and tremolo at di erent speeds. C11.P53

• Pitch contours: various shapes of successive tones and sub-tone in ections. C11.P54

• Modal shapes: cumulative pitch images with distinct interval constellations. C11.P55

• Harmonic shapes: single chords and/or several chords in succession. C11.P56

• Spectral shapes, both stationary and dynamic timbral features. C11.P57

• Rhythmic shapes: patterns of sound onsets and durations. C11.P58

• Textural shapes: wefts of sound events. C11.P59

• Articulation shapes: accents, staccato, legato, and various particular acts. C11.P60

• Timing shapes: accelerando, ritardando, and various uctuations. C11.P61

The crucial property of shapes here is that they are distributed, i.e. re ect the temporal unfolding of both C11.P62
sound and body motion, and are also inherently holistic in that they are perceivable ‘all at once’. This means
that shapes are non-symbolic in the sense of what Schae er called ‘concrete’, as opposed to that which is
symbol-oriented and called ‘abstract’ (1966). In short, shapes are manifestations of musical instants as
overview images of that which unfolds at the meso timescale, typically in the 0.3–3 seconds duration range.
Also, images of shape can work across modalities, and hence be the basis for ‘sound-motion objects’, i.e.
salient multimodal meso timescale fragments of music.
Sound-Motion Objects C11.S7

The elements of intermittency and of shape converge in what can be called ‘sound-motion objects’. Based C11.P63
on Pierre Schae er’s concept of sound objects, sound-motion objects are, as already mentioned, basically a
multimodal extension of Schae er’s idea of sound objects by also including sensations of body motion in
musical experience (Godøy 2006). And similar to Schae er’s notion of objects, the object view of sound and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
body motion here is based on the idea that sequentially unfolding features may be kept in memory as
cumulative retrospective and/or anticipatory images, typically including the dynamic, pitch-related, and
timbre-related envelopes, as well as the trajectories of the sound-producing and/or sound-accompanying
body motion.

As an example, think of the sound of someone saying wah-wah where there is an envelope of timbre and C11.P64
dynamics, as well as an envelope of the mouth changing shape. The sound is continuous from the start to
p. 210 the end of the fragment, and so is the motion of the mouth, from the beginning when the mouth starts to
open through to the end when the mouth is closing. And in the course of this trajectory, the mouth motion
results in the distinct spectral shape of the wah-wah sound (as we know, on a trumpet or on a trombone, a
similar wah-wah sound would be made by opening and closing a mute). This means that we now can
combine previous insights on gestalt-based object coherence in auditory perception theory (e.g. Bregman
1990) with insights from gestalt-based motor control theory (Klapp and Jagacinski 2011), and also
supplement these insights with ndings on the e ciency of object attention (De Freitas et al. 2014).

The main element here is the combination of intermittency in motor control, i.e. of the serial ballistic C11.P65
control (Loram et al. 2014), with the slower and noise-prone motor system of our bodies, so that
intermittent impulses set the musculoskeletal system in motion, in turn producing a continuous sound
output. This means that each sound-motion object is triggered by an impulse, and hence each sound-
motion object may be seen as focused on one instant—something I have tried to document with data on
velocity reversals in sound-producing body motion, e.g. with the impact-points in piano performance
(Godøy 2013).

We may thus think of sound-motion objects as multimodal, just as much objects of body motion as of C11.P66
sound. This also means that body motion coherence may become a schema for sound coherence, hence the
idea that musical gestalts are also body motion gestalts. Furthermore, we can hypothesize that sensations of
musical instants are linked to sensation of intermittent control in body motion, that the serial ballistic
control scheme results in a series of body motion chunks that, due to their very nature of being controlled
intermittently, may be perceived as coherent entities and as continuous within the limits of their duration.
Stated di erently, the idea here is that sensations of the musical instant are closely linked with sensations
of intermittency in human body motion, and that the impulsive triggering of sound-producing body motion
chunks may also contribute to the subjective sensations of musical instants.

Conclusions and Prospects C11.S8

The main point of trying to understand musical instants is to reconcile continuity and discontinuity in C11.P67
musical experience. This can arguably be done by demonstrating that sound and body motion are both
continuous and coherent at the meso timescale of sound-motion objects, yet that these sound-motion
objects are driven by intermittent impulses of motor control and e ort. In summary, sound-motion objects
are:

• preprogrammed holistic entities; C11.P68

• impulse-driven entities; C11.P69


• centred on key-postures; C11.P70

• internally coherent and continuous in time as multimodal sound and body motion gestalts. C11.P71

p. 211 Furthermore, musical instants, based on the intermittency of both motor control and motor e ort, may C11.P72
provide insights on:

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
• The human body’s strategy for generating fast and accurate music-related motion in the not-so-fast C11.P73
motor control system of our bodies by using anticipatory, feedforward, control impulses.

• The optimization of skilled music-related body motion by intermittent e ort, cf. experts’ energy C11.P74
reduction by minimizing muscle contraction times and exploiting conservation of momentum.

Insights into the workings of musical instants could have practical applications in music performance, C11.P75
improvisation, and composition by:

• thinking all kinds of rhythmical patterns, including syncopations and polyrhythmic patterns, as single C11.P76
impulse driven body motion patterns (cf. Klapp et al. 1998);

• practising ornaments as pre-programmed sound-motion objects driven by impulses, e.g., in fast drum C11.P77
set lls (Godøy et al. 2017) or fast violin trills (Godøy et al. 2016);

• conceiving of all kinds of textures as piecewise impulse-driven, e.g., tremolo, trills, arpeggios, dance C11.P78
patterns, etc. as impulse-driven sound-motion objects (Godøy 2018);

• providing tools for studying high-level salient features such as style, aesthetics, a ect, and sense of C11.P79
motion, as well as by focusing on the holistic features of meso timescale sound-motion objects (Godøy
2018).

Needless to say, we have signi cant challenges here, rst of all in documenting the existence of C11.P80
intermittent control in music-related body motion. The cited research (Karniel 2013; Loram et al. 2014;
Sakaguchi et al. 2015) has used a variety of more indirect methods for this, and other directly signal-based
methods also seem promising (Inoue and Sakaguchi 2015). In my own work and that of colleagues, we have
tried to look at features of body motion data, in particular of acceleration and velocity reversals, as well as
data on trajectory coherence and smoothness, for traces of intermittency.

Presently we are supplementing motion capture data with EMG data in order to see what muscle activation C11.P81
patterns can tell us about intermittency in both control and e ort, as well as to develop an impulse-
response type model of intermittent control in sound-producing motion (cf. the work reported in Berio et al.
2017; Plamondon et al. 2013). Furthermore, sound features can tell us something about intermittency and
coherence, and we believe that various tools from music information retrieval (MIR) can be useful in
extracting perceptually salient shape and intermittency data from sound les—data that can be correlated
with the corresponding sound-producing and sound-accompanying body motion data.

p. 212 All these research e orts could contribute to our understanding of musical instants as a crucial element in C11.P82
musical experience, including its potential practical applications in performance, improvisation,
composition, and sound design. The overall goal here is to demonstrate how music is based on a series of
subjectively perceived and conceived instants, as well as how these instants fuse into experiences of larger-
scale musical works.
Notes

1. Our ability to provide fast, seemingly instantaneous judgements or answers (e.g. read words on a large billboard or C11.N1
calculate 2 + 2 =?), is called ʻSystem 1ʼ by Daniel Kahneman, and what he calls ʻSystem 2ʼ denotes our ability to solve more
complex tasks (e.g. fill out a tax form or calculate 17 24 =?), tasks that for most people require more extended and step-by-
step reasoning.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
2. Schae er came to see the a inity of his method with phenomenology, but he commented that he and his colleagues, with C11.N2
this pragmatic phonograph loop origin and use of reduced listening, were initially actually practising phenomenology
without knowing it (Schae er 1966: 262).
References C11.S9

Aoki, H., Tsukahara, R., and Yabe, K. (1989). E ects of pre-motion electromyographic silent period on dynamic force exertion C11.P83
during a rapid ballistic movement in man. European Journal of Applied Physiology 58(4): 426–432.
Google Scholar WorldCat

Berio, D., Akten, M., Leymarie, F. F., Grierson, M., and Plamondon, R. (2017). Calligraphic stylisation learning with a C11.P84

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
physiologically plausible model of movement and recurrent neural networks. Proceedings of the 4th International Conference on
Movement Computing. https://doi.org/10.1145/3077981.3078049
Google Scholar Google Preview WorldCat COPAC

Bizley, J. K., and Cohen, Y. E. (2013). The what, where and how of auditory-object perception. Nature Reviews Neuroscience 14: C11.P85
693–707.
Google Scholar WorldCat

Bregman, A. (1990). Auditory scene analysis. MIT Press. C11.P86


Google Scholar Google Preview WorldCat COPAC

Chemin, B., Mouraux, A., and Nozaradan, S. (2014). Body movement selectively shapes the neural representation of musical C11.P87
rhythms. Psychological Science 25(12): 2147–2159.
Google Scholar WorldCat

De Freitas, J., Liverence, B. M., and Scholl, B. J. (2014). Attentional rhythm: A temporal analogue of object-based attention. C11.P89
Journal of Experimental Psychology 143(1): 71–76.
Google Scholar WorldCat

Eitan, Z., and Granot, R. Y. (2008). Growing oranges on Mozartʼs apple tree: ʻInner formʼ and aesthetic judgment. Music Perception C11.P90
25(5): 397–417.
Google Scholar WorldCat

Galantucci, B., Fowler, C. A., and Turvey, M. T. (2006). The motor theory of speech perception reviewed. Psychonomic Bulletin and C11.P91
Review 13(3): 361–377.
Google Scholar WorldCat

Gaver, W. W. (1993). What in the world do we hear? An ecological approach to auditory event perception. Ecological Psychology C11.P92
5(1): 1–29.
Google Scholar WorldCat

Gawthorp, P., Loram, I., Lakie, M., and Golle, H. (2011). Intermittent control: a computational theory of human control. Biological C11.P93
Cybernetics 104: 31–51.
Google Scholar WorldCat

Gjerdingen, R., and Perrott, D. (2008). Scanning the dial: The rapid recognition of music genres. Journal of New Music Research C11.P94
37(2): 93–100.
Google Scholar WorldCat

p. 213 Godøy, R. I. (1997). Formalization and epistemology. Scandinavian University Press. C11.P95
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2001). Imagined action, excitation, and resonance. In R. I. Godøy and H. Jorgensen (eds), Musical imagery, 239–252. C11.P96
Swets & Zeitlinger.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2003). Motor-mimetic music cognition. Leonardo 36(4): 317–319. C11.P97


Google Scholar WorldCat

Godøy, R. I. (2006). Gestural-sonorous objects: embodied extensions of Schae erʼs conceptual apparatus. Organised Sound C11.P98
11(2): 149–157.
Google Scholar WorldCat

Godøy, R. I. (2008). Reflections on chunking in music. In A. Schneider (ed.), Systematic and comparative musicology: concepts, C11.P99
methods, findings, 117–132. Peter Lang.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2010a). Gestural a ordances of musical sound. In R. I. Godøy and M. Leman (eds), Musical gestures: Sound, C11.P100
movement, and meaning, 103–125. Routledge.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2010b). Images of sonic objects. Organised Sound 15(1): 54–62. C11.P101
Google Scholar WorldCat

Godøy, R. I. (2010c). Thinking now-points in music-related movement. In R. Bader, C. Neuhaus, and U. Morgenstern (eds), C11.P102
Concepts, experiments, and fieldwork: Studies in systematic musicology and ethnomusicology, 245–260. Peter Lang.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2011). Sound-action awareness in music. In D. Clarke and E. Clarke (eds), Music and consciousness, 231–243. Oxford C11.P103
University Press.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2013). Quantal elements in musical experience. In R Bader (ed.), Sound–perception–performance: Current research in C11.P104
systematic musicology, vol. 1, 113–128. Springer.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2014). Understanding coarticulation in musical experience. In M. Aramaki, M. Derrien, R. Kronland-Martinet, and C11.P105
S. Ystad (eds), Sound, music, and motion: Lecture notes in computer science, 535–547. Springer.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2017). Key-postures, trajectories and sonic shapes. In D. Leech-Wilkinson and H. Prior (eds), Music and shape, 4–29. C11.P106
Oxford: Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I. (2018). Sonic object cognition. In R. Bader (ed.), Springer handbook of systematic musicology, 761–777. Springer C11.P107
Nature.
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I., and Leman, M. (eds) (2010). Musical gestures: Sound, movement, and meaning. Routledge. C11.P108
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I., Song, M., and Dahl, S. (2017). Exploring sound-motion textures in drum set performance. In Proceedings of the 14th C11.P109
Sound and Music Computing Conference, Aalto University.
http://smc2017.aalto.fi/media/materials/proceedings/SMC17_p145.pdf
Google Scholar Google Preview WorldCat COPAC

Godøy, R. I., Song, M., Nymoen, K., Romarheim, M. H., and Jensenius, A. R. (2016). Exploring sound-motion similarity in musical C11.P110
experience. Journal of New Music Research 45(3): 210–222.
Google Scholar WorldCat

Gra on, S. T., and Hamilton, A. F. (2007) Evidence for a distributed hierarchy of action representation in the brain. Human C11.P111
Movement Science 26: 590–616.
Google Scholar WorldCat

Gri iths, T. D. and Warren, J. D. (2004). What is an auditory object? Nature Reviews Neuroscience 5(11): 887–92. C11.P112
Google Scholar WorldCat

Haken, H., Kelso, J., and Bunz, H. (1985). A theoretical model of phase transitions in human hand movements. Biological C11.P113
Cybernetics 51(5): 347–356.
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
Hardcastle, W., and Hewlett, N. (eds) (1999). Coarticulation: Theory, data and techniques. Cambridge University Press. C11.P114
Google Scholar Google Preview WorldCat COPAC

Hindemith, P. (2000). A composer's world: Horizons and limitations. Schott.


Google Scholar Google Preview WorldCat COPAC

Hogan, N., and Sternad, D. (2007). On rhythmic and discrete movements: Reflections, definitions and implications for motor C11.P115
control. Experimental Brain Research 181: 13–30.
Google Scholar WorldCat

Husserl, E. (1982). Ideas pertaining to a pure phenomenological philosophy: First book. Kluwer Academic. C11.P116
Google Scholar Google Preview WorldCat COPAC

p. 214 Husserl, E. (1991). On the phenomenology of the consciousness of internal time, 1893 –1917. (trans. J. B. Brough). Kluwer C11.P117
Academic.
Google Scholar Google Preview WorldCat COPAC

Inoue, Y., and Sakaguchi, Y. (2015). A wavelet-based method for extracting intermittent discontinuities observed in human motor C11.P118
behavior. Neural Networks 62: 91–101.
Google Scholar WorldCat

Kahneman, D. (2011). Thinking, fast and slow. Farrar, Straus and Giroux. C11.P119
Google Scholar Google Preview WorldCat COPAC

Karniel, A. (2013). The minimum transition hypothesis for intermittent hierarchical motor control. Frontiers in Computational C11.P120
Neuroscience 7. https://doi.org/10.3389/fncom.2013.00012
Google Scholar WorldCat

Klapp, S. T., and Jagacinski, R. J. (2011). Gestalt principles in the control of motor action. Psychological Bulletin 137(3): 443–462. C11.P121
Google Scholar WorldCat

Klapp, S. T., Nelson, J. M., and Jagacinski, R. J. (1998). Can people tap concurrent bimanual rhythms independently? Journal of C11.P122
Motor Behavior 30(4): 301–322.
Google Scholar WorldCat

Loram, I. D., van de Kamp, C., Lakie, M., Gollee, H., and Gawthrop, P. J. (2014). Does the motor system need intermittent control? C11.P123
Exercise and Sport Science Review 42(3): 117–125.
Google Scholar WorldCat

Michon, J. (1978). The making of the present: A tutorial review. In J. Requin (ed.), Attention and Performance VII, 89–111. Erlbaum. C11.P124
Google Scholar Google Preview WorldCat COPAC

Petitot, J. (1990). Forme. In Encyclopædia Universalis. Encyclopædia Universalis. C11.P125


Google Scholar Google Preview WorldCat COPAC

Plamondon, R., OʼReilly, C., Rémi, C., and Duval, T. (2013). The lognormal handwriter: Learning, performing, and declining. C11.P126
Frontiers in Psychology 4(945). https://doi.org/10.3389/fpsyg.2013.00945
Google Scholar WorldCat

Pöppel, E. (1997). A hierarchical model of time perception. Trends in Cognitive Science 1(2): 56–61. C11.P127
Google Scholar WorldCat

Rosenbaum, D. A. (2010). Human motor control (2nd edn). Elsevier/Academic Press. C11.P128
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469713 by National Science & Technology Library user on 26 May 2023
Rosenbaum, D. A. (2017). Knowing hands: The cognitive psychology of manual control. Cambridge University Press. C11.P129
Google Scholar Google Preview WorldCat COPAC

Rosenbaum, D., Cohen, R. G., Jax, S. A., Weiss, D. J., and van der Wel, R. (2007). The problem of serial order in behavior: Lashleyʼs C11.P130
legacy. Human Movement Science 26(4): 525–554.
Google Scholar WorldCat

Sakaguchi, Y., Tanaka, M., and Inoue, Y. (2015). Adaptive intermittent control: A computational model explaining motor C11.P131
intermittency observed in human behavior. Neural Networks 67: 92–109. http://dx.doi.org/10.1016/j.neunet.2015.03.012.
Google Scholar WorldCat

Schae er, P. (1966). Traité des objets musicaux. Seuil. C11.P132


Google Scholar Google Preview WorldCat COPAC

Suied, C., Agus, T. R., Thorpe, S. J., Mesgarani, N., and Pressnitzer, D. (2014). Auditory gist: Recognition of very short sounds from C11.P133
timbre cues. Journal of the Acoustical Society of America 135(3): 1380–1391.
Google Scholar WorldCat

Thom, R. (1983). Paraboles et catastrophes. Flammarion. C11.P134


Google Scholar Google Preview WorldCat COPAC

Xenakis, I. (1992). Formalized Music (rev. edn). Pendragon Press. C11.P135


Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
CHAPTER

12 C12
Cross-Modality and Embodiment of Tempo and Timing

Renee Timmers

https://doi.org/10.1093/oxfordhb/9780190947279.013.10 Pages 215–C12.P104


Published: 08 December 2021

Abstract
This chapter explores the insights that research into cross-modal correspondences and multisensory
integration o er to our understanding and investigation of tempo and timing in music performance.
As tempo and timing are generated through action, actions and sensory modalities are coupled in
performance and form a multimodal unit of intention. This coupled intention is likely to demonstrate
characteristics of cross-modal correspondences, linking movement and sound. Testable properties
predictions are o ered by research into cross-modal correspondences that have so far mainly found
con rmation in controlled perceptual experiments. For example, fast tempo is predicted to be linked to
smaller movement that is higher in space. Con rmation in the context of performance is complicated
by interacting associations with intentions related to e.g. dynamics and energy, which can be
addressed through appropriate experimental manipulation. This avenue of research highlights the
close association between action and cross-modality, conceiving action as a source of cross-modal
correspondences as well as indicating the cross-modal basis of actions. For timing and tempo
concepts, action and cross-modality o er concrete and embodied modalities of expression.

Keywords: crossmodal correspondences, tempo, timing, music performance, embodied cognition


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

DO musical tempo and time primarily exist within the auditory domain or are they co-expressed in other C12.P1
modalities, in particular in the movements and gestures of performers? The aim of this chapter is to bring
together two strands of research that seem to have had hitherto little contact or communication. This
concerns on the one hand research on (expressive) timing and movement in performance, including in
ensemble contexts, and on the other hand research on cross-modal correspondences with auditory
1
characteristics, in particular with tempo and time. Research on movement in performance has investigated
the role of movements in the production and perception of music performance, paying special attention to
the use of ancillary (non-sound-producing) movements. This research has, for example, investigated how
movements relate to musical structure or to auditory aspects of the performance, including relationships
with phrasing, tempo, and intensity (MacRitchie et al. 2013; Timmers et al. 2006; Wanderley et al. 2005).
Additionally, research has investigated how movements are involved in communication between
performers, as well as between performers and audiences, demonstrating the relevance of body movement
for the communication of emotion, expressive intentions, and the provision of temporal cues (see e.g.
Bishop and Goebl 2018; Davidson 1993; 2005; Vines et al. 2006; Vuoskoski et al. 2014). In this chapter, I
argue that many of these investigations could bene t from the inclusion of hypotheses related to the way in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
which movements and auditory characteristics may be associated, as predicted by studies examining cross-
modal correspondences of auditory characteristics. Similarly, studies examining cross-modal
correspondences could bene t from a closer investigation of how multiple modalities are engaged in the
production of sound, time, and tempo.

As studies of cross-modal correspondences generally examine correspondences within carefully controlled C12.P2
p. 216 perceptual experiments (see Spence 2011 for a review), it is necessary to bring together evidence found
through variations of isolated variables and perceptual judgements and observations made in multivariate,
ecological performance contexts. In this chapter, I begin to bridge this gap by formulating a framework
tackling some of the complexities involved in relating experimental results to ecological performance
contexts. However, before formulating this framework and looking into speci c examples, I consider in
more detail relevant aspects of marking time and tempo in performance, and discuss predictions from the
literature on cross-modal correspondences of tempo and timing.

Continuous Movements, Discrete Musical Events, and Grouping C12.S1

Hierarchies

In music performance, continuous movements generate discrete events, and, in turn, such discrete events C12.P3
may generate impressions of continuous movement (Gjerdingen 1994). Discrete pitch events can in most
instances be de ned within music performances, even for instruments that have continuous pitch, such as
the voice, Theremin, or trombone, or when instruments transition between pitch events with a glissando or
portamento. Figure 12.1 shows a fundamental pitch trajectory of a tenor singing the opening notes of The
Swan (1886) by Saint-Saëns. In this example, continuous and gradual changes in the vibration of the vocal
folds give rise to ve discrete tones that move in pitch: a high opening tone, followed by two lower tones, a
return to a relatively high tone, and another descent to a lower tone.
Figure 12.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Fundamental pitch trajectory of the opening tones of Saint-Saënsʼ The Swan (1886), sung by a tenor at a moderate tempo C12.F1
(reproduced from Desain et al. 1999).

p. 217 Onsets of these discrete pitch events and the intervals between them are particularly salient for the C12.P4
perception of time and tempo. In Figure 12.1, the time interval between successive tone onsets uctuates.
Nevertheless, these time intervals are categorized as equal, rounded to an approximate 1:1 ratio. Such
periodically spaced tone onsets form the basis of tempo perception, and uctuations in metrically placed
tone onsets will be perceived as changes in local tempo (Dixon et al. 2006). Fluctuations in the interval
between perceived tone onset and o set and their relationship to inter-onset intervals are perceived as
changes in articulation (Repp 1995). While the exact perceptual processes involved are complex, and fall
outside the scope of this chapter, this example illustrates some fundamental aspects of markings of musical
time, in particular how tone onsets underpin the perception of rhythm and metre, and how small changes in
duration may generate local tempo variation.

During performance, we may assume that performers time the onsets of tones. They also time the tone C12.P5
o sets, but possibly with less precision and consistency. Consequently, we may distinguish between
movements that immediately precede tone onsets, movements that occur during the duration of the tone,
and movements that follow event o set. However, researchers of time perception and production in music
would argue that the notion of musical events used in this description is too simple. Onsets that occur
rapidly after each other are grouped and perceived as combined events. Also, in the timing of music, events
at the level of the beat with a certain minimal duration may be timed directly, while faster events forming a
motive or rhythmic gure are timed as a group (e.g. Clarke 1987). Godøy (2014) emphasizes the bene ts of
such coarticulation of events in music performance, as it facilitates fast and smooth movement production.
How exactly notes are timed may depend on the performer, working within constraints o ered by
perception and skilled motor control. In any case, it may be more informative to distinguish movements
leading up to a group of tones, during the group, and following the group, if we are interested in deducing
the expressive intention, including the marking of time and tempo.

Indeed, a considerable proportion of research into communicative movements of performers has looked at C12.P6
larger timespans, going beyond the level of single notes and beats to examine movement trajectories within
bars, phrases, and groups of phrases (MacRitchie et al. 2013; Vines et al. 2006). As Figure 12.2 illustrates,
periodicities within movements of, in this case, the head of a pianist may vary considerably depending on
the musical section of a piece. In this example, relatively fast periodicities had similar durations as two-bar
subphrases, while slower periodic movements spanned several subphrases (see Figure 12.2). This raises the
question of whether movement is more accurately perceived and more meaningful at a slower timescale
than auditory events. There is some evidence that this is the case. For example, performers coordinate
body-sway which spans multiple beats to facilitate synchronization at the beat level (Keller and Appel
2010). Nevertheless, movement-to-sound and movement-to-movement entrainment (synchronization) is
meaningful at a range of speeds and a range of temporal levels, as exempli ed by coordinated dance,
including at the level of the tactus (see e.g. van Noorden and Moelants 1999 for an overview of tempi in
p. 218 dance music). Given that the intentional timing of performers happens at the level of beats and groups of
beats and events, we may expect temporal information to be decoded in movements at a hierarchy of levels
too.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Figure 12.2

Le –right and backward–forward movement of the head of a pianist in three performances by the same pianist of Scriabinʼs C12.F2
Étude No. 12 in D# minor Op. 8 (1894) (reproduced from Timmers et al. 2006; see also Camurri et al. 2003).

Cross-modal Correspondences with Time and Tempo C12.S2

The previous section examined the temporal relationship between movement and musical timing C12.P7
characteristics, and the temporal levels at which we may expect to see alignment. Central to this section is
the question of which movement properties may correlate with musical time and tempo. Should we consider
p. 219 the amount of movement displacement (measuring how much of the body moves and how far), the speed
of movements, the position, acceleration patterns, spatial direction, shape? Research has shown that
when conductors indicate the timing of beats, peaks in acceleration preceding a reversal in direction of
movement are the main markers to which musicians synchronize (Luck and Sloboda 2008; 2009). In other
words, it is the movement kinematics, in particular acceleration, that are communicative, irrespective of the
spatial location of the particular movements. Analogously, in a study of violin-violin, violin-piano and
piano-piano duos in which one or the other performer was assigned to be the leader, the timing and
periodicity of head acceleration peaks were found to best align with temporal characteristics of the music in
the context of a leader and follower playing a duo and the leader using cuing-in movements to start the
performance (Bishop and Goebl 2018).
Nevertheless, expressive intentions, including those related to tempo and timing, may be communicated C12.P8
through a broader range of movement features than acceleration only. In analyses of dance and performers’
movements, Camurri and colleagues use measures of the amount of movement, which relates both to the
size of the circumference of the movement and the distance travelled per time unit, contraction, or
expansion of the body, the amount of upward movement, the length and direction of movement
trajectories, as well as kinematic cues related to velocity and acceleration patterns (Camurri et al. 2003).
From these, further characteristics may be calculated such as movement uency, entropy, or impulsiveness

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
(see also Broughton and Davidson 2016). Some of these characteristics have been con rmed to in uence
observers’ perceptions of performance. For example, the length of visual gestures in uences the perceived
duration of tones in the context of marimba performances (Schutz and Lipscomb 2007). Moreover, even if
the visual duration is not di erent from the audible duration, the width of a gesture in uences the perceived
duration of a sung tone (Cai et al. 2013). At a more general level, it has been shown that the amount of
movement in uences observers’ evaluations of the expressivity of a performance, which is otherwise
related to the amount of uctuations in tempo and dynamics (Davidson 1993; Kendall and Carterette 1990;
Vuoskoski et al. 2014), and that such properties as speed, smoothness, and range of movement inform
observers’ evaluations of the emotion intended in a performance (Dahl and Friberg 2007).

As mentioned above, a central point of this chapter is to argue that cross-modal correspondences should be C12.P9
consulted when interpreting and aiming to understand relationships between sonic properties and
visual/movement characteristics. Whilst this may seem obvious, this argument has received surprisingly
little attention in music performance research. An important strength of research on cross-modal
correspondences is that it o ers a number of predictions related to the types of mappings or analogies that
we can expect between movement, physical and visual properties, and sound. Although potentially a vast
number of properties may be associated with each other, we may start with a smaller set of dimensions that
have been shown to relate systematically (Walker et al. 2012). This set includes associations between small
size, fast tempo, high pitch, light weight, thin, sharp, and bright objects. On the other end of the spectrum,
p. 220 large size and slow tempo are further associated with low pitch, heavy weight, thick, round, and dark
objects. Some of these correspondences have a statistical origin, such as the association between large
objects and low pitch, while others may have a structural-neurological origin, such as the association
between high pitch and bright light (Spence 2011). Other correspondences seem without a statistical,
neurological, or linguistic explanation and Walker et al. (2012) argue that these may arise through cross-
talk between correspondences (see also Walker 2016; Wallmark 2019). For example, an unexpected
association is that observers expect a brightly coloured snooker ball to be lighter in weight than a darkly
coloured snooker ball (Walker and Walker 2012). Experiments showing interference of judgments of one
dimension by variations in the other dimension demonstrate that these mappings are not just conventional
associations, but internalized intuitions that in uence our perception (Casasanto and Boroditsky 2008;
Timmers and Li 2016). Applying these ndings to performers’ movements, we may expect that movements
associated with faster tempi and shorter time intervals will not only be faster in speed, but also smaller in
size or circumference, lighter in weight, as well as possibly higher in space and more angular in shape. By
contrast, slower sounds will be played with larger and rounder movements that are possibly heavier in
weight and lower in position. These associations can be said to be used in a communicative sense if they, for
example, facilitate co-performers’ synchronization, if they are also present in unconstrained anticipatory
movements, or if they are more pronounced in collaborative music-making contexts.

A complication occurs for these predictions if we consider in addition the role of intensity (or sound C12.P10
amplitude, as an indication of dynamics). Fast tempo and high intensity are associated, as are slow tempo
and soft intensity (for a review, see Eitan 2017). High intensity has partially opposite associations from fast
tempo: high intensity is heavy in weight and large in size rather than light and small, which are associated
with soft sounds. These opposite mappings may interfere if fast tempo and high intensity or energy are
communicated; then movements may be expected to be larger, heavier, as well as higher in space.
Furthermore, for some features such as pitch, associations with dynamic characteristics (rising vs. falling
pitch) are di erent from associations with static characteristics (high and low pitch). For example, while
rising pitch may connote increases in size, high pitch is small (Eitan et al. 2014). Furthermore, high pitch
has been shown to be associated with fast tempo and low pitch with slow tempo, but the same was not found
for rising versus falling pitch (Tamir-Ostrover and Eitan 2015). For tempo, contrasting associations may
occur with respect to weight: through its association with high energy, increasing tempo may be associated
with increases in force and weight rather than becoming lighter.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Cross-modal correspondences with and between movements and sounds are not restricted to this set of C12.P11
correspondences. We may, for example, also draw a connection to emotional expression as several
correspondences have emotional implications such as dark vs. bright, low vs. high pitch, and low vs. high
energy, which relate to emotional valence and arousal (Eitan and Timmers 2010; Eitan et al. 2017). However,
p. 221 for now, the set displayed in Table 12.1 will make up our main hypotheses.

Table 12.1. Cross-modal correspondences with slow and fast tempo, and so and loud sounds C12.T1

Modality Slow tempo Fast tempo So Loud

Movement Slow Fast Slow Fast

Size Large Small Small Large

Spatial location Low High Low High

Weight Heavy Light Light Heavy

Shape Round Sharp Round Sharp

Colour Dark Light Low High

Pitch Low High Low High

Intensity So Loud

Energy Low High Low High


Developing the framework using related findings from the literature C12.S3

What evidence can be found in existing studies for the ways in which movements mark time or C12.P12
communicate tempo? First, in considering the ways in which movements have been observed to mark time,
we can examine periodicities in movement patterns and their correspondence to rhythmical aspects of the
music, as well as the phase relationship between musical structure and movement patterns. Starting with

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
the rst, in an exploratory study of advanced pianists’ upper-body movement patterns, periodicities in
movements were found to relate to a two-bar phrase structure of the performed music, which in some
instances additionally showed a hierarchical organization across phrases (MacRitchie et al. 2013). Although
the movement patterns were highly individual to pianists, autocorrelation analyses indicated similarities in
the underlying periodicities of the individual movements. The communicative potential of these movements
was shown by the demonstration that presentation of video as well as audio recordings of performances to
participants improved their indications of phrase structure (MacRitchie et al. 2013). Other studies have
similarly shown that audiovisual presentation of performances in uences the indication of perceived
phrase structure of the music in addition to in uencing other perceived properties of the musical
communication, such as musical tension, emotion, or expressivity (Davidson 1993; Timmers et al. 2006;
Vines et al. 2006).

Periodic movement at the level of sub-phrases or phrases is a commonly observed feature in piano C12.P13
performances (MacRitchie et al. 2013), and has also been observed in other instruments (Wanderley et al.
p. 222 2005). It highlights the larger timespans with which music is timed and coordinated. Typical for musical
timing is its hierarchical nature: periodicities at the beat level cluster into half-bars and bars, bars into
subphrases, and subphrases into phrases, and so on. Such nested temporal relationships can also be
observed in movements associated with music, whether in the context of spontaneous and stylized dance
movements to music (Leman and Naveda 2010; Toiviainen et al. 2010), or in the movements of performers.
Movement periodicities on several levels are simultaneously present and may be actualized in di erent
areas of the body or di erent movement directions. For example, foot-taps and head-bops at beat level may
combine with shoulder-sway at a slower pace (Leman and Naveda 2010).

Periodicity analysis illustrates coupling between cross-sensory temporal modalities, which may not be a C12.P14
surprise as movement is directly involved in the generation of temporal structure in the auditory domain,
and indeed we may expect that this not only concerns the production of tones but also the expression of
larger-scale temporal units (Clarke 1999). Nevertheless, the relationship is not a straightforward one. The
phase relationship between sound and movement is more variable than one might expect if movement is
assumed to generate sound, and movement expressions are highly idiosyncratic, even in ensemble
performance contexts, where sounds are closely coordinated (Glowinski et al. 2013). For example, phase-
locking is considerably more exible and less tight between changes in movement direction and sub-phrase
grouping boundaries than is the case for troughs in local tempo and dynamics (Timmers et al. 2006). Bishop
and Goebl (2018) found that peaks in movement acceleration are more strongly coupled to beat onsets than
peaks (or troughs) in movement direction and that this coupling was more strongly present for assigned
leaders in a duo than followers. Even so, the coupling was variable, as indicated by the proportion of
acceleration peaks in the movement of the head that aligned with beats, which was calculated to be 57% for
head acceleration patterns. Bowing hand acceleration patterns in violinists showed a larger alignment
proportion of 67%, which still seems lower than expected given the close relationship between bowing and
tone production.

To understand the role of movement in music performance and relationships between characteristics of the C12.P15
movement, sound, and musical structure, researchers have often drawn from literature on gestures in
speech. Body movements show closest parallels with gesticulations, where speech is simultaneously
present, rather than iconic gestures, that may serve a communicative purpose on their own (McNeill 2005).
In gesticulations, verbal and nonverbal behaviours are expressions of a communicative intention in
channel- (or modality-) speci c manners that are in support of each other (e.g. McNeill 2005). We may
assume that a repertoire of gestures and sounds form the basis of and shape expressive communication and
that this repertoire is developed through experience and training in speci c contexts. Indeed, it is likely that
the expressive intention is multimodal, as is its actualization. Movements are not just employed in the
service of produced sounds, but sounds and movements are elements of a multimodal expressive form in
which sounds and movement are intended in coordination, although shaped by modality speci c-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
constraints. For example, the intention to communicate tension vs. relaxation will shape sounds as well as
p. 223 movements and is likely to be conceptualized in terms of movement patterns and intended sounds (Li
and Timmers 2020). This is in line with studies that show the relevance and complementarity of visual
information for the perception of music (Schutz 2008), and interpretations of body movement and
performed sound as components within a complex dynamic system (Demos et al. 2014).

As argued above, cross-modal correspondences can be expected to be at the heart of a meaningful and C12.P16
complementary relationship between sounds and movements. A complication for the examination of cross-
modal correspondences in music performance is that even though relationships between sound and motion
are frequently objects of investigation, studies have not explicitly examined these relationships from the
perspective of cross-modal correspondences. To nd supporting or disproving evidence, we need to rely on
instances of reported relationships in studies that were conducted with a di erent objective in mind.

The rst relationship is relatively straightforward: e ectors move faster when producing faster tempi (top C12.P17
row of Table 12.1). However, this does not only need to concern the e ectors producing the sounds such as
the ngers. Other body parts may similarly be moving faster as we found when correlating peak velocities of
head movements in two-bar phrases with peaks in performed tempo in these two-bar phrases (Timmers et
2
al. 2006). This correlation was, however, only modest when compared to the relationship between
movement velocity and key velocity or intensity of the music, which were found to have a strong
relationship. While these correlations were observed in the context of solo performance, they may be
observed for communicative purposes in ensemble contexts too. Indeed, we found that the duration of the
bow of the gesture preceding the rst onset of the music was correlated with the subsequent tempo of the
performances of a string quartet. This correlation was apparent for the assumed leader of the group, the
rst violin (Timmers et al. 2014). Similarly, Bishop and Goebl (2018) found that the duration of head and
bowing gestures of violinists performing in a duo varied linearly as predicted with indicated tempo, in
particular for the assigned leader of the duo.

These studies did not investigate other characteristics of the movements, such as the size or circumference C12.P18
of the movement, the shape, or the spatial location. Relationships between shape and tempo can be
observed in studies investigating conducting in which the speed of movement and the radius of curvature
are inversely related (Luck and Sloboda 2009), corresponding to relatively sharper movements.
Furthermore, relationships between spatial height and tempo have been observed in two studies
investigating the kinematics of pianists’ nger movements at di erent tempi. Palmer and Dalla Bella
(2004) observed that at faster tempi ngers were raised higher. This nding was replicated in Dalla Bella
and Palmer (2011). The raising of the ngers corresponded with a larger distance of movement and greater
movement speeds, even if normalized according to performed tempo. The use of raised ngers, further
increasing the speed and distance of movements, is in contrast to kinematic predictions, that would predict
more e cient and smaller movement trajectories at a faster tempo. The raising of ngers may be related to
the intention to play fast, and the high activity and high vertical position associated with it (see Table 12.1).

p. 224 On the other hand, faster tempo may also be associated with smaller movement, corresponding to less C12.P19
highly raised ngers, and indeed this has a kinematic advantage for speed production (Haken et al. 1985).
What the speci c correlation may be is likely to depend on the expressive intention of the performer. If a
performance is intended to be high in energy, the tempo will be faster, the movements larger rather than
smaller, and the movements possibly jumpier, higher, and more angular. In other instances, a performance
may be fast, lightweight, and small in movement, even though it may also travel greater distances across
the piano keyboard. The power of the consideration of cross-modal correspondences is then not to generate
absolute predictions about relationships between movement and temporal aspects of music that will be true
irrespective of context; rather, it provides a framework of logical heuristics that performers may employ for
expressive and creative purposes (Leech-Wilkinson and Prior 2014).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Cross-modal Markings of Time and Tempo: Some New Data C12.S4

As a nal consideration of the relationship between movement characteristics and performed tempo and C12.P20
timing, I will consider some new data collected in the context of a study that investigated the role of seeing
the movements of a co-performer for the ability to synchronize with that performer (Timmers et al. 2020).
In this study, movements of the performing hand of a pre-recorded performer were tracked and visualized
as an animation presented to musician participants. The task of the participants was to perform as closely as
possible in synchrony with several pre-recorded performances (performed by a single pianist), whilst
hearing the recording, hearing and seeing a video recording, or hearing and seeing the animation. For the
current analysis, we can investigate the relationship between temporal and movement characteristics in the
pre-recorded performances and investigate the potential usefulness of seeing these movements to the
participants in their aim to synchronize with the recording. The recordings were of performances of the
left-hand part of four simple pieces by Bartók, in which the left and right hand are playing the same
material an octave or sixth apart. The left-hand performances were performed without metronome and had
some, but not extensive, variations in overall and local tempo. Video recordings were made of the
movements of the left hand and a small part of the lower arm. Movement of the hand was tracked from the
video recordings focussing on the movement of the metacarpophalangeal joint of the index nger of the left
hand (for further details, see Timmers et al. 2020).

For the purposes of this chapter, we will examine correlations between local duration measured as inter- C12.P21
onset-intervals (IOI) and the vertical position of the hand, as well as the distance of hand displacement.
Vertical position and amount of displacement were measured at di erent temporal locations with respect to
note onsets and note o sets. Speci cally, movement characteristics were summarized for four temporal
p. 225 positions: in the frame immediately preceding note onset (–1), in the frame coinciding with the note
onset ( 0), in the frames between note onset and o set (1), in the frame coinciding with the note o set (2),
and in the frames in between note o set and the frame preceding the next note onset (3). Table 12.2 presents
the observed correlations for quarter notes and half notes, as these were the two note durations within the
pieces. For comparison, it also shows analogous correlations between key velocity, hand position, and hand
displacement.
Table 12.2. Spearmanʼs Rho correlations between height of hand (le ) or hand displacement (right), local duration (IOI), and C12.T2
intensity (key velocity) for four temporal positions: immediately preceding note onset (–1), at note onset ( 0), during note
duration (1), at note o set (2), and in between note o set and note onset (3). Correlations are separately calculated for quarter
notes and half notes. (N=190 for quarter notes, N=62 for half notes.)

Position Height of hand (vertical position) Hand displacement (absolute)

IOI Key velocity IOI Key velocity

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Quarter Half Quarter Half Quarter Half Quarter Half
notes notes notes notes notes notes notes notes

–1 .149* .306* .140 .294* .119 .176 .404** .353**

0 .208** .322* .176* .244 –.023 .286* .405** .369**

1 .325** .424** .017 .256* .263** –.122 .224** .055

2 .369** .303* .045 .244 .010 .157 .285** .120

3 .389** .169 .155* .237 –.292** .030 .390** .029

* p < .05,

** p < .01

In contrast to the ndings of Palmer and Dalla Bella (2004; Dalla Bella and Palmer 2011), these C12.P22
performances showed a positive relationship between the height of the hand and variations in the quarter
note and half note IOIs, which corresponds to a negative association between height and tempo. As
discussed earlier, this is in line with predictions from movement kinematics, which would predict smaller
movements, moving less high with respect to the piano keyboard for faster tempi. The variations in tempo
are smaller in this study than in the study of Palmer and Dalla Bella, and the performer was not instructed to
perform at a faster or slower tempo. These di erences in context may be responsible for the di erence in
results where for this performance associations between small movement size and faster tempo dominate,
while for discrete tempo instructions intentions to perform at a certain tempo in uence the displayed
kinematics.

The degree of hand displacement is most strongly correlated to key velocity, which is a measure of the C12.P23
p. 226 intensity of the performance. Hand displacement shows a weaker relationship to IOI, and one that varies
in direction. For note onset and sustain, the correlation is positive: more movement for longer durations.
However, after note release, the relationship is negative: with longer durations, the hand moves less. This
negative correlation further strengthens to rs = –.449, if we correct the amount of displacement for the
duration of this moment. This negative correlation is what we expect based on cross-modal
correspondences: slow tempi (longer durations) are associated with less movement. The positive
relationship may be related to the greater possibility for movement during longer durations. This variability
exempli es the need for systematic control of the various constraints on movements in order to be able to
draw solid inferences.

These movements and variations in the qualities of movements operate at a fast temporal scale and it can be C12.P24
questioned whether these variations have any communicative function. They may have an expressive
function in the sense that they are part of the communicative intention of the performing pianist, but this
does not mean that they are also informative to co-performers or an audience. As mentioned above, one of
the objectives of the study of Timmers et al. (2020) was to investigate whether co-performers used the
visual information of the performer and whether this contributed positively to the ability to synchronize
with the pre-recorded performance. To investigate this, 25 musically trained participants were asked to
synchronize with the recorded performers under three audiovisual conditions as explained before. Four
levels of pianistic expertise were distinguished: 0 (non-pianists), 1 (amateur pianists), 2 (semi-professional
pianists), and 3 (professional pianists). Asynchronization was measured as the standard deviation of the
di erences in onset timing between the pre-recorded performances and performances of the participants.
Larger standard deviations of onset timing asynchronies were interpreted as indicative of less consistent
and therefore less successful synchronization.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
While the full experiment included conditions with Transcranial Magnetic Stimulation (TMS) of particular C12.P25
brain areas, in this analysis we focus on the data for the TMS sham condition in which the brain was not
actually stimulated. Only synchronization data for particular target notes in the music were considered as
these notes show relatively precise timing and therefore results are less blurred by noisiness in the
production data. An ANOVA was conducted to test the e ects of audiovisual condition, note duration
(quarter note or half note), and pianistic expertise on asynchronization (standard deviation of onset timing
di erences). This ANOVA showed a signi cant interaction between three main e ects: the relationship
between pianistic expertise and asynchronization varied with note duration and audiovisual condition (F(2,
40) = 4.00, p = .026, r = .41). Table 12.3 shows the correlations between the asynchronization measure and
pianistic expertise for each of the six conditions, and shows stronger, signi cant negative correlations
between expertise and asynchronization for the conditions that include visual information compared to the
audio-only condition. This indicates that the presence of visual information increases expertise e ects on
synchronization ability. Indeed, high-level experts on the same instrument may bene t from visual
information, while visual movement information may not help or may even disadvantage performers who
p. 227 are less expert or have high expertise on a di erent instrument.

Table 12.3. Correlations between level of pianistic expertise and asynchronization (standard deviation of onset timing C12.T3
di erences) in di erent experimental conditions

Note duration Audio only Audio–video Audio–animation

Quarter notes –.092 –.586** –.180

Half notes –.365 –.216 –.531**

** p < .01, df = 24

The full study argues and presents evidence that positive uses of visual movement information is related to C12.P26
the ability to internally simulate the actions of the co-performer, implying the involvement of mirror-
neurons or high levels of perception-action coupling (Keller et al. 2014), in addition to the use of this
information as a source for temporal cuing. The link from simulating or modelling the co-performers’
actions to accurate synchronization is a cross-modal translation, grounded in multimodal performance
experience.
General Discussion of Theoretical Framework C12.S5

The starting point for this chapter was the question of whether musical tempo and time primarily exist C12.P27
within the auditory domain or whether they are co-expressed in other modalities, in particular in the
movements and gestures of performers. This question was addressed by investigating links between
temporal characteristics of movement and sound in performance, and exploring how notions of cross-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
modal correspondences may enhance our understanding and insight into these links. It was argued that
cross-modal correspondences are central to music performance, following from the understanding of
performance as the simultaneous expression of an intention in movement and sound. This further
highlights the relevance of perception–action coupling and embodiment for our understanding of
performed time and tempo. By providing a framework for the interpretation of relationships between sound
and movement, predictions were formulated that can be empirically tested in experimental investigations. A
few examples were examined, which provided some preliminary evidence, but also indicated the
complexities of testing the reality of cross-modal manifestations of time and tempo characteristics in
performers’ movements, and the need to test predictions experimentally.

Several aspects central to the argued relationship between movement and temporal aspects of music are C12.P28
p. 228 well established. In particular, the automaticity of perceptual associations between certain properties in
di erent sense modalities has been demonstrated rigorously using various experimental paradigms often
including some kind of speeded response task that requires inhibition of a response to particular
information, and sometimes including the measurement of a neurological electrophysiological response in
addition to behavioural measures (see e.g. Bien et al. 2012). Furthermore, the notion that musical properties
induce cross-modal associations is well established. Studies have asked participants to move while music is
playing, which may be through whole body movement, moving the arm to draw on a projected screen,
drawing on paper, or choosing how an animated character would move (Eitan and Granot 2008; Kohn and
Eitan 2016; Küssner et al. 2014). These studies highlight the multiplicity and multidimensionality of
possible mappings. Mappings are not predetermined or xed, given a certain sound property, but are
probabilistic, contextual, and individual. Furthermore, relationships are not always bi-directional and may
be asymmetrical, varying with the speci c pole of a bipolar continuum that is presented (see Eitan 2017;
Eitan and Timmers 2010). This means that while there is solid evidence that these correspondences have a
strong presence and can in uence perception, the speci c manifestation of correspondences and their
interpretation has a fair degree of uncertainty.

The strength of this chapter is therefore not to make xed predictions about the intended meaning of C12.P29
movements with particular properties, or about what movements should be present in the context of certain
musical characteristics. Nevertheless, it does o er predictions as in Table 12.1, which should help to
constrain the range of expected cross-modal manifestations. Primarily, I argue that relationships between
sound, movement and cross-modal correspondences should be investigated in tandem, including in the
context of music performance, but also when interpreting dance, as together they o er a promising
theoretical framework. Not only are cross-modal correspondences likely to manifest in the context of
performers’ movements, actions may be an important source for cross-modal correspondences to
originate, as demonstrated in Timmers and Li (2016). Even though this chapter is not the rst to argue for a
‘web of auditory-motor-cognitive mappings’, as Kohn and Eitan formulate it (2016: 40), a central role of
embodiment for cross-modal correspondences is not explicitly recognized in existing review literature
(Eitan 2017; Spence 2011; Walker 2016). Acknowledging the central role of body–environment interaction
for cognition (Rowlands 2010; Varela et al. 1991), it seems self-evident that cross-modal correspondences
also originate from action, which is in line with neurophysiological research that has shown the motoric
encoding of auditory phenomena, in both speech and music (Anumanchipalli et al. 2019; Zatorre et al.
2007).
In this chapter, I have outlined several directions in which this framework can take us. First of all, the C12.P30
intention to perform is multimodal: intentions consist of the actions and movements that we use and the
intended sounds and sequences of sounds that we aim to produce. One is not just subsidiary to the other.
Secondly, intentions and realizations are shaped by modality-speci c characteristics, such as the
requirement for precise onset timing in the auditory domain, and the continuous kinematics of movements.
p. 229 Thirdly, movements and sounds cohere together through a meaningful relationship. For the reasons
mentioned under the rst and second premise, this may be a considerably more complex relationship than

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
an exact one-to-one mapping. Nevertheless, we may assume that movement and sound cohere together,
not just in time and in structure, but also in underlying characteristics. Previous research has argued
something similar in the context of emotional expression or degree of musical expressiveness (e.g.
Davidson 1993; Vuoskoski et al. 2014), when investigating the e ect of seeing a performer on the perception
of dynamics or timing (Vuoskoski et al. 2016) or when comparing perception of structure in dance and
music (Krumhansl and Schenk 1997). Nevertheless, it is generally not embraced fully or systematically,
which brings me to my nal and most speculative premise that properties that can be expressed cross-
modally have prioritized meaning in the context of music-making. As music-making has a strong
grounding in physical performance and dance, it is likely to meaningfully express properties related to these
physical motions and experiences. This may concern emotional properties, but is certainly not restricted to
it. Based on cross-modal research, we may expect physical properties to feature including size, mass,
height, speed, energy, force, brightness, smoothness, warmth, roundedness, distance, tension, and torsion.

Given the complexities of relationships observed in ecological contexts without speci c instructions and C12.P31
experimental manipulations, as was the case for the examples referred to in this chapter, an important
direction for future research will be to target certain relationships more explicitly in the investigation. This
may, for example, be through instructing an individual to perform in certain ways (e.g. fast or slow, short or
long, far or near) and observe how movement properties vary. It may also be through perceptual
experiments that vary movement and sonic properties simultaneously, translating paradigms such as in
Casasanto and Boroditsky (2008) for time and distance to the context of music performance (as in Cai et al.
2013). However, taking a more practical approach, it could be particularly useful to investigate how cross-
modal concepts feature or may feed into music education and support musical thinking. For example, Li and
Timmers (2020) found that bodily and cross-modal metaphors were intentionally used to produce di erent
timbres in piano performance. This intention concerned changes in bodily movements as much as or even
more than changes in the produced sounds. Furthermore, it would be of interest to consider the
relationships between music and dance from the perspective of cross-modal correspondences, examining
how movement and associated concepts may feed into the understanding of musical characteristics and vice
versa, and using correspondences to interpret the way dancers relate their movements to music and the two
channels create meaning.

Taking this discussion back to the perception and communication of time and tempo: sounds are excellent C12.P32
markers and manipulators of time and tempo, including the creation of an illusion of continuous
acceleration, the presence of a hierarchy of larger and smaller timespans, and the ability for time to seem
unending and come to a standstill. While this chapter is primarily concerned with concrete ways in which
p. 230 movements and temporal characteristics of music are related in the context of music performance,
cross-modal correspondences are likely to contribute to these broader examples too. Psychologists building
on gestalt psychology often alluded to visual perception, when explaining parallel phenomena in auditory
perception, such as the illusion of continuation (Gjerdingen 1994). Theories have used kinematic properties
of movement as part of explanatory frameworks for perceptual constraints on tempo and time (Drake et al.
2000; Todd 1995; van Noorden and Moelants 1999), and studies including those of dance have shown the
reciprocal relationship between tempo perception and movement characteristics such as the vigorousness
of the dance (London et al. 2016). Analogously, the lack of perceived movement and direction in sound may
be a means for time to become unmarked and eternal.
For the communication between performers and between performers and their audience, cross-modal C12.P33
correspondences are relevant in their simultaneous presence in sound and movement, in the
communication of intentions whilst performing, and in the meaningful shaping of time, tempo, duration,
and intensity. For example, the perception of musical characteristics such as groove, swing, laid-back, and
funk may be related to how musicians and dancers move as much as the audible manner of timing. To
interpret such timing, we need to examine parallels with movement and its properties, rather than focus
primarily on temporal and auditory properties. In this context, it is important to note that understanding

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
the timing and rhythm of a genre is supported by learning how to move to it, as has been argued in the
context of interpreting Bulgarian uneven meters (Moelants 2006; Rice 1994).

While this perspective is strongly embodied, it does not negate the power of concepts and rich associations C12.P34
that may go beyond what is physically possible, creating the ability for imaginary transcendence through
music: music induces notions of space, movement, size, time, and tempo that may go beyond the everyday,
whether this concerns the energy and size suggested by a large ensemble performance or the tempo and
movement variation of a virtuoso soloist. On the other hand, for successful ensemble performance, it may
be su cient for communication to remain fairly concrete through a multimodal alignment of sounds and
actions. The framework proposed here predicts that aligning the associated properties of these sounds and
actions should help to enhance synchronization and the sharing of performance goals.

Acknowledgements C12.S6

Part of this research was made possible through funding from the Leverhulme Trust (IAF-2015-013). The C12.P35
data presented in this chapter was part of a larger data set collected during a research visit at the MARCS
institute of Western Sydney University. The larger study was conducted in collaboration with Peter Keller,
Jennifer Macritchie, Siobhan Schabrun, Tribikram Thapa, and Manuel Varlet. I would like to thank Emily
Payne, Zohar Eitan, Dorothy Ker, and Rolf Inge Godøy for helpful feedback on the chapter.

p. 231
Notes
1. Cross-modal correspondences refer to characteristics in di erent modalities, such as auditory and visual, that are C12.N1
perceived as corresponding, or belonging together. For example, short duration is associated with small size (Rojczyk
2011), and fast tempo with high energy, fast movement and visual brightness (Collier and Hubbard 1998; Eitan and Granot
2006).

2. Note that in the cited study, inter-onset intervals were used as measure of variation in speed instead of tempo, and C12.N2
therefore a negative correlation was found between local duration (IOI) and movement velocity.
References C12.S7

Anumanchipalli, G. K., Chartier, J., and Chang, E. F. (2019). Speech synthesis from neural decoding of spoken sentences. Nature C12.P36
568: 493–498.
Google Scholar WorldCat

Bien, N., ten Oever, S., Goebel, R., and Sack, A. T. (2012). The sound of size: crossmodal binding in pitch-size synaesthesia: a C12.P37

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
combined TMS, EEG and psychophysics study. NeuroImage 59(1): 663–672.
Google Scholar WorldCat

Bishop, L., and Goebl, W. (2018). Beating time: How ensemble musiciansʼ cueing gestures communicate beat position and C12.P38
tempo. Psychology of Music 46(1): 84–106.
Google Scholar WorldCat

Broughton, M. C., and Davidson, J. W. (2016). An expressive bodily movement repertoire for marimba performance, revealed C12.P39
through observersʼ Laban e ort-shape analyses, and allied musical features: Two case studies. Frontiers in Psychology 7: 1211.
https://doi.org/10.3389/fpsyg.2016.01211
Google Scholar WorldCat

Cai, Z. G., Connell, L., and Holler, J. (2013). Time does not flow without language: Spatial distance a ects temporal duration C12.P40
regardless of movement or direction. Psychonomic Bulletin and Review 20(5): 973–980.
Google Scholar WorldCat

Camurri, A., Mazzarino, B., Ricchetti, M., Timmers, R., and Volpe, G. (2003). Multimodal analysis of expressive gesture in music C12.P41
and dance performances. In A. Camurri and G. Volpe (eds), International gesture workshop, 20–39. Springer.
Google Scholar Google Preview WorldCat COPAC

Casasanto, D., and Boroditsky, L. (2008). Time in the mind: Using space to think about time. Cognition 106: 579–593. C12.P42
Google Scholar WorldCat

Collier, W. G., and Hubbard, T. L. (1998). Judgments of happiness, brightness, speed and tempo change of auditory stimuli C12.P43
varying in pitch and tempo. Psychomusicology 17(1–2): 36.
Google Scholar WorldCat

Clarke, E. F. (1987). Levels of structure in the organization of musical time. Contemporary Music Review 2(1): 211–238. C12.P44
Google Scholar WorldCat

Clarke, E. F. (1999). Rhythm and timing in music. In D. Deutsch (ed.), The psychology of music (2nd edn), 473–500. Academic C12.P45
Press.
Google Scholar Google Preview WorldCat COPAC

Dahl, S., and Friberg, A. (2007). Visual perception of expressiveness in musiciansʼ body movements. Music Perception 24(5): 433– C12.P46
454.
Google Scholar WorldCat

Dalla Bella, S., and Palmer, C. (2011). Rate e ects on timing, key velocity, and finger kinematics in piano performance. PloS One C12.P47
6(6): e20518. https://doi.org/10.1371/journal.pone.0020518
Google Scholar WorldCat

Davidson, J. W. (1993). Visual perception of performance manner in the movements of solo musicians. Psychology of Music 21(2): C12.P48
103–113.
Google Scholar WorldCat

Davidson, J. W. (2005). Bodily communication in musical performance. In D. Miell, R. A. MacDonald, and D. J. Hargreaves (eds), C12.P49
Musical communication, 215–237. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

p. 232 Desain, P., Honing, H., Aarts, R., and Timmers, R. (1999). Rhythmic aspects of vibrato. In P. Desain and L. Windsor (eds), Rhythm C12.P50
perception and production, 203–216. Swets & Zeitlinger.
Google Scholar Google Preview WorldCat COPAC

Demos, A. P., Cha in, R., and Kant, V. (2014). Toward a dynamical theory of body movement in musical performance. Frontiers in C12.P51

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Psychology 5: 477. https://doi.org/10.3389/fpsyg.2014.00477
Google Scholar WorldCat

Dixon, S., Goebl, W., and Cambouropoulos, E. (2006). Perceptual smoothness of tempo in expressively performed music. Music C12.P52
Perception 23(3): 195–214.
Google Scholar WorldCat

Drake, C., Jones, M. R., and Baruch, C. (2000). The development of rhythmic attending in auditory sequences: attunement, C12.P53
referent period, focal attending. Cognition 77(3): 251–288.
Google Scholar WorldCat

Eitan, Z. (2017). Musical connections: Crossmodal correspondences. In R. Ashley and R. Timmers (eds), The Routledge companion C12.P54
to music cognition, 213–224. Routledge.
Google Scholar Google Preview WorldCat COPAC

Eitan, Z., and Granot, R. Y. (2006). How music moves: Musical parameters and listeners images of motion. Music Perception 23(3): C12.P55
221–248.
Google Scholar WorldCat

Eitan, Z., and Granot, R. Y. (2008). Growing oranges on Mozart's apple tree: ʻInner formʼ and aesthetic judgment. Music Perception C12.P56
25(5): 397–418.
Google Scholar WorldCat

Eitan, Z., and Timmers, R. (2010). Beethovenʼs last piano sonata and those who follow crocodiles: Cross-domain mappings of C12.P57
auditory pitch in a musical context. Cognition 114(3): 405–422.
Google Scholar WorldCat

Eitan, Z., Schupak, A., Gotler, A., and Marks, L. E. (2014). Lower pitch is larger, yet falling pitches shrink. Experimental Psychology C12.P58
61(4): 273–284.
Google Scholar WorldCat

Eitan, Z., Timmers, R., and Adler, M. (2017). Cross-modal correspondences and a ect in a Schubert song. In D. Leech-Wilkinson C12.P59
and H. Prior (eds), Music and shape, 58–86. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Gjerdingen, R. O. (1994). Apparent motion in music? Music Perception 11(4): 335–370. C12.P60
Google Scholar WorldCat

Glowinski, D., Mancini, M., Cowie, R., Camurri, A., Chiorri, C., and Doherty, C. (2013). The movements made by performers in a C12.P61
skilled quartet: A distinctive pattern, and the function that it serves. Frontiers in Psychology 4: 841.
https://doi.org/10.3389/fpsyg.2013.00841
Google Scholar WorldCat

Godøy, R. I. (2014). Understanding coarticulation in musical experience. In M. Aramaki, O. Derrien, R. Kronland-Martinet, and C12.P62
S. Ystad (eds), Sound, music, and motion. CMMR 2013. Springer. https://doi.org/10.1007/978-3-319-12976-1_32

Haken, H., Kelso, J. S., and Bunz, H. (1985). A theoretical model of phase transitions in human hand movements. Biological C12.P63
Cybernetics 51(5): 347–356.
Google Scholar WorldCat

Keller, P. E., and Appel, M. (2010). Individual di erences, auditory imagery, and the coordination of body movements and sounds C12.P64
in musical ensembles. Music Perception 28(1): 27–46.
Google Scholar WorldCat

Keller, P. E., Novembre, G., and Hove, M. J. (2014). Rhythm in joint action: Psychological and neurophysiological mechanisms for C12.P65

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
real-time interpersonal coordination. Philosophical Transactions of the Royal Society B: Biological Sciences 369(1658): 20130394.
Google Scholar WorldCat

Kendall, R. A., and Carterette, E. C. (1990). The communication of musical expression. Music Perception 8(2): 129–163. C12.P66
Google Scholar WorldCat

Kohn, D., and Eitan, Z. (2016). Moving music: Correspondences of musical parameters and movement dimensions in childrenʼs C12.P67
motion and verbal responses. Music Perception: An Interdisciplinary Journal 34(1): 40–55.
Google Scholar WorldCat

Krumhansl, C. L., and Schenck, D. L. (1997). Can dance reflect the structural and expressive qualities of music? A perceptual C12.P68
experiment on Balanchineʼs choreography of Mozartʼs Divertimento No. 15. Musicae Scientiae 1(1): 63–85.
Google Scholar WorldCat

p. 233 Küssner, M. B., Tidhar, D., Prior, H. M., and Leech-Wilkinson, D. (2014). Musicians are more consistent: Gestural cross-modal C12.P69
mappings of pitch, loudness and tempo in real-time. Frontiers in Psychology 5: 789. https://doi.org/10.3389/fpsyg.2014.00789
Google Scholar WorldCat

Leech-Wilkinson, D., and Prior, H. M. (2014). Heuristics for expressive performance. In D. Fabian, R. Timmers, and E. Schubert C12.P70
(eds), Expressiveness in music performance: Empirical approaches across styles and cultures, 34–57. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Leman, M., and Naveda, L. (2010). Basic gestures as spatiotemporal reference frames for repetitive dance/music patterns in C12.P71
Samba and Charleston. Music Perception 28(1): 71–91.
Google Scholar WorldCat

Li, S., and Timmers, R. (2020). Exploring pianistsʼ embodied concepts of piano timbre: An interview study. Journal of New Music C12.P72
Research 49(5): 477–492.
Google Scholar WorldCat

London, J., Burger, B., Thompson, M., and Toiviainen, P. (2016). Speed on the dance floor: auditory and visual cues for musical C12.P73
tempo. Acta Psychologica 164: 70–80.
Google Scholar WorldCat

Luck, G., and Sloboda, J. (2008). Exploring the spatio-temporal properties of simple conducting gestures using a synchronization C12.P74
task. Music Perception 25(3): 225–239.
Google Scholar WorldCat

Luck, G., and Sloboda, J. A. (2009). Spatio-temporal cues for visually mediated synchronization. Music Perception 26(5): 465–473. C12.P75
Google Scholar WorldCat

MacRitchie, J., Buck, B., and Bailey, N. J. (2013). Inferring musical structure through bodily gestures. Musicae Scientiae 17(1): 86– C12.P76
108.
Google Scholar WorldCat

McNeill, D. (2005). Gesture and thought. University of Chicago Press. C12.P77


Google Scholar Google Preview WorldCat COPAC
Moelants, D. (2006). Perception and performance of aksak metres. Musicae Scientiae 10(2): 147–172. C12.P78
Google Scholar WorldCat

Palmer, C., and Dalla Bella, S. (2004). Movement amplitude and tempo change in piano performance. Journal of the Acoustical C12.P79
Society of America 115(5): 2590.
Google Scholar WorldCat

Repp, B. H. (1995). Acoustics, perception, and production of legato articulation on a digital piano. Journal of the Acoustical C12.P80

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Society of America 97(6): 3862–3874.
Google Scholar WorldCat

Rice, T. (1994). May it fill your soul: Experiencing Bulgarian music. University of Chicago Press. C12.P81
Google Scholar Google Preview WorldCat COPAC

Rojczyk, A. (2011). Sound symbolism in vowels: Vowel quality, duration and pitch in sound-to-size correspondence. Poznań C12.P82
Studies in Contemporary Linguistics 47: 602–615.
Google Scholar WorldCat

Rowlands, M. (2010). The new science of the mind: From extended mind to embodied phenomenology. MIT Press. C12.P83
Google Scholar Google Preview WorldCat COPAC

Schutz, M. (2008). Seeing music? What musicians need to know about vision. Empirical Musicology Review 3: 83–108. C12.P84
Google Scholar WorldCat

Schutz, M., and Lipscomb, S. (2007). Hearing gestures, seeing music: Vision influences perceived tone duration. Perception 36(6): C12.P85
888–897.
Google Scholar WorldCat

Spence, C. (2011). Crossmodal correspondences: A tutorial review. Attention, Perception, and Psychophysics 73(4): 971–995. C12.P86
Google Scholar WorldCat

Tamir-Ostrover, H., and Eitan, Z. (2015). Higher is faster: Pitch register and tempo preferences. Music Perception 33(2): 179–198. C12.P87
Google Scholar WorldCat

Timmers, R., Endo, S., Bradbury, A., and Wing, A. M. (2014). Synchronization and leadership in string quartet performance: a case C12.P88
study of auditory and visual cues. Frontiers in Psychology 5: 645. https://doi.org/10.3389/fpsyg.2014.00645
Google Scholar WorldCat

Timmers, R., and Li, S. (2016). Representation of pitch in horizontal space and its dependence on musical and instrumental C12.P89
experience. Psychomusicology 26(2): 139–148.
Google Scholar WorldCat

Timmers, R., MacRitchie, J., Schabrun, S. M., Thapa, T., Varlet, M., and Keller, P. E. (2020). Neural multi-modal integration C12.P90
underlying synchronization with a co-performer in music: influences of motor expertise and visual information. Neuroscience
Letters 721: 134803. https://doi.org/10.1016/j.neulet.2020.134803
Google Scholar WorldCat

p. 234 Timmers, R., Marolt, M., Camurri, A., and Volpe, G. (2006). Listenersʼ emotional engagement with performances of a Scriabin C12.P91
étude: An explorative case study. Psychology of Music 34(4): 481–510.
Google Scholar WorldCat

Todd, N. P. M. (1995). The kinematics of musical expression. Journal of the Acoustical Society of America 97(3): 1940–1949. C12.P92
Google Scholar WorldCat

Toiviainen, P., Luck, G., and Thompson, M. R. (2010). Embodied meter: Hierarchical eigenmodes in music-induced movement. C12.P93
Music Perception 28(1): 59–70.
Google Scholar WorldCat

van Noorden, L., and Moelants, D. (1999). Resonance in the perception of musical pulse. Journal of New Music Research 28(1): C12.P94
43–66.
Google Scholar WorldCat

Varela, F. J., Thompson, E., and Rosch, E. (1991). The embodied mind: Cognitive science and human experience. MIT Press. C12.P95

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469805 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Vines, B. W., Krumhansl, C. L., Wanderley, M. M., and Levitin, D. J. (2006). Cross-modal interactions in the perception of musical C12.P96
performance. Cognition 101(1): 80–113.
Google Scholar WorldCat

Vuoskoski, J. K., Thompson, M. R., Clarke, E. F., and Spence, C. (2014). Crossmodal interactions in the perception of expressivity C12.P97
in musical performance. Attention, Perception, and Psychophysics 76(2): 591–604.
Google Scholar WorldCat

Vuoskoski, J. K., Thompson, M. R., Spence, C., and Clarke, E. F. (2016). Interaction of sight and sound in the perception and C12.P98
experience of musical performance. Music Perception 33(4): 457–471.
Google Scholar WorldCat

Walker, P. (2016). Cross-sensory correspondences: A theoretical framework and their relevance to music. Psychomusicology C12.P99
26(2): 103–116. https://doi.org/10.1037/pmu0000130
Google Scholar WorldCat

Walker, L., Walker, P., and Francis, B. (2012). A common scheme for cross-sensory correspondences across stimulus domains. C12.P100
Perception 41(10): 1186–1192.
Google Scholar WorldCat

Walker, P., and Walker, L. (2012). Size–brightness correspondence: Crosstalk and congruity among dimensions of connotative C12.P101
meaning. Attention, Perception, and Psychophysics 74(6): 1226–1240.
Google Scholar WorldCat

Wallmark, Z. (2019). Semantic crosstalk in timbre perception. Music and Science 2: 1–8. C12.P102
https://doi.org/10.1177/2059204319846617
Google Scholar WorldCat

Wanderley, M. M., Vines, B. W., Middleton, N., McKay, C., and Hatch, W. (2005). The musical significance of clarinetistsʼ ancillary C12.P103
gestures: An exploration of the field. Journal of New Music Research 34(1): 97–113.
Google Scholar WorldCat

Zatorre, R. J., Chen, J. L., and Penhune, V. B. (2007). When the brain plays music: auditory–motor interactions in music C12.P104
perception and production. Nature Reviews Neuroscience 8(7): 547–558.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
CHAPTER

13 The Mind Is a DJ: Rhythmic Entrainment in C13

Beatmatching and Embodied Temporal Processing 


Maria A. G. Witek

https://doi.org/10.1093/oxfordhb/9780190947279.013.13 Pages 235–252


Published: 08 December 2021

Abstract
In music, rhythmic entrainment occurs when the attention and body movements of listeners, dancers
and musicians become synchronized with the beat. This synchronization occurs due to the
mechanisms of phase and period correction. Here, I describe what happens to these mechanisms
during beatmatching—a central skill in DJing that involves synchronizing the beats of two records on a
set of turntables. Via the enactivist approach to the embodied mind, I argue that beatmatching a ords
a di erent form of entrainment that requires more conscious control of and embodied
operationalization of temporal error correction, and thus provides a vivid model of the embodied
distribution of rhythmic entrainment.

Keywords: DJing, beatmatching, temporal processing, entrainment, embodiment, enactivism


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

I’m a scientist before I’m anything C13.P1

Grandmaster Flash C13.P2

WHEN musicians groove together, listeners tap along, and dancers synchronize their bodies to the beat of C13.P3
the music, their attention and movements are driven by the process of entrainment. Research into beat
perception and sensorimotor synchronization has shown us that such entrainment relies on the
mechanisms of temporal error correction. A cellist who lags behind will move their bow faster to correct for
any time di erence between their playing and the others in the quartet. By contrast, a DJ cannot rely on this
direct sensorimotor relationship between movement and timing when mixing records together on a set of
turntables. Instead, they must manipulate the records and adjust the controllers on the decks to align the
tempi of the tracks—a process known as beatmatching. What happens to the process of entrainment during
beatmatching in DJing? What constraints are put on the mechanisms of temporal error correction, and what
does it mean for our understanding of the mind that these corrections happen on the turntables, as opposed
to inside the head of the DJ?

In this chapter, I review the temporal perception mechanisms involved in sensorimotor synchronization C13.P4
and rhythmic entrainment and analyse what happens to these processes when a DJ is beatmatching. Via the
theory of enactivism—in which the coupling between agents and their environments (including other
p. 236 agents) forms the basis for embodied life and mind—I argue that beatmatching presents a form of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
entrainment that has yet to be considered in music psychology and philosophy of mind. In beatmatching,
temporal error correction must be consciously controlled via the skilful manipulation of the records and the
turntables. In this way, the beatmatching DJ o ers an unusually vivid example of the embodied distribution
of rhythmic entrainment and its underlying temporal correction mechanisms.

Rhythmic Entrainment and the Beat C13.S1

The beat is at the centre of dance music. It is a regularly occurring and usually isochronous pattern of salient C13.P5
moments in music. Di erent forms of dance music emphasize the beat in di erent ways. In electronic dance
music, such as disco, house, and techno, the beat is commonly marked directly with a relatively sonically
dominating kick-drum (known as ‘four on the oor’). In other genres, such as hip-hop and jazz, the beat is
often partly implied by the rhythmic structure, through o -beat phrasing and interlocking syncopation.

Our ability to perceive a steady beat lies at the most basic level of rhythm perception and production. C13.P6
Without this ability, our experience of music would lose any sense of metric structure, complex structures
like syncopations and polyrhythms would lose their a ective feel, and our ability to play or dance in time to
music would be thwarted. Beat perception is de ned as the perception of temporal regularity in a rhythmic
pattern, an ability that is largely innate and universal (Honing 2012) but which can be improved through
musical training (Repp and Su 2013). It has recently become clear that it is not a uniquely human capacity,
following a string of studies investigating beat perception and production in non-human animals such as
parrots, sea lions, and monkeys (Cook et al. 2013; Gámez et al. 2018; Patel et al. 2009), but it is in humans
that this capacity is most developed.

While the cognitive mechanisms are still debated (Grahn 2012; Jones 2009), the theory of entrainment is C13.P7
currently the most widely cited model that explains how we perceive and synchronize to a beat. In the
dynamic attending approach (Barnes and Jones 2000; Jones and Boltz 1989; Large and Jones 1999), rhythms
or beat patterns are thought of in terms of oscillations, and entrainment is the process by which an
independent and self-sustaining oscillator comes into contact with and becomes driven by another
independent and self-sustaining oscillator, causing them to synchronize. The three main properties that
de ne an oscillator are its phase, period, and amplitude. The phase is the temporal position of an oscillator,
while the period (or frequency) is its speed. The amplitude refers to the strength of the oscillator’s signal. In
order for two unsynchronized oscillators to become fully entrained, either or both must adapt in phase
and/or period. Oscillators that have the same period but di erent phase will move at the same speed, but
one will be o set by a value known as the asynchrony. Oscillators that have the same phase but di erent
periods can start o in time but gradually desynchronize, causing cumulating asynchronies as one oscillator
moves faster than the other. The exception to this is when the two periods are in a simple ratio relationship
p. 237 to one another. For example, a 2:1 ratio leads every phase of the rst oscillator to align with every other
phase of the second oscillator.

Entrainment occurs naturally in various physical (e.g. pendulum clocks) and biological systems (e.g. C13.P8
circadian sleep–wake cycles) (Clayton et al. 2004). In music and in rhythmic coordination more generally,
entrainment is thought to explain how humans are able to perceive and synchronize with a rhythmic beat,
as when tapping their feet or dancing along. From this perspective, the beat and the listener (or musician, or
dancer) are viewed as oscillators, and it is the entrainment of their attentional processes to the music that
drives the perception of the beat (Large and Kolen 1994). Entrainment can occur between an individual and
an external rhythm, and when that external rhythm is generated by another individual, this is called
interpersonal or social entrainment. (Phillips-Silver et al. 2010). There can also be self-entrainment, as
when an individual entrains to a self-generated or imagined rhythm. It is important to distinguish between
entrainment and mere synchrony. If two clocks placed in opposite ends of a house happen to be ticking
perfectly in time, they are synchronized, but no entrainment has occurred. There needs to be some level of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
adaptation in phase and/or period for there to be true entrainment.

Temporal expectations play a crucial role in our entraining to a beat. When a repetitive beat is perceived, our C13.P9
attentional oscillations form expectancies about where the future beats will occur (Barnes and Jones 2000).
This is partly re ected by the tendency to slightly anticipate the beat when synchronizing movements to it,
by tapping approximately 50 ms before the stimulus onset (Repp and Su 2013). Temporal expectations are
also central to metre perception, where hierarchically organized and recursively subdivided layers of
regularly occurring beats or pulses lead to perceived patterns of alternating perceptual salience (Honing
2012). For example, when two layers (or oscillations) are related in a 2:1 ratio, as illustrated above, the
alignment of phases at the higher level is given more salience than the individual phases at the lower level.
The more salient the metric position, the stronger the expectations are that rhythmic events will occur at
that position.

Beat perception and rhythmic entrainment are usually studied through measurements of the coordination C13.P10
of rhythmic movement with a beat, a process known as ‘sensorimotor synchronization’. Finger-tapping
studies are particularly common, as they o er a simple and intuitive way to measure synchronized
movements that produce discrete temporal events (Repp and Su 2013). They o er a simple model of a wide
range of musical activities in which musicians synchronize their playing on an instrument with an external
musical beat, either played by another musician or a recording. Motion-capture is increasingly being used
to record continuously synchronized movements of a greater variety of body parts and musical contexts
(Toiviainen et al. 2009; Witek et al. 2017). While most humans e ortlessly synchronize their movements to
a regular beat, their motor activity always exhibits some degree of temporal variability, i.e. asynchrony. In
addition to simple motor noise—i.e. variability resulting from uctuations in the motor system—
individuals express synchronization variability that re ects how strongly a beat is perceived, with stronger
p. 238 beat perception leading to less variable tapping (Wing and Kristo erson 1973). In order to maximize
synchrony, constant monitoring of this asynchrony is needed. In entrainment terms, the phase and period
of the movement oscillation must be continuously corrected according to the perceived asynchrony, or
error, between the taps and the beat.

When tapping in synchrony with a steady isochronous sequence, such as a metronome, the placement of C13.P11
each tap is adjusted by phase correction. In order to maintain synchrony with the beat, the tapper has to
estimate the asynchronies between their taps and the metronomic beats, and temporal adjustments are
made that correct these asynchronies. Psychological experiments have shown that phase correction is
largely automatic (Repp and Keller 2004). However, while no awareness is required, phase correction can be
voluntarily controlled. Experiments in which participants are told to ignore perturbations when tapping to
otherwise regular sequences show that while the phase correction response cannot be entirely suppressed,
it can be signi cantly reduced (Repp 2002a; 2002b). This conscious ability to correct for phase asynchrony
appears to depend on the magnitude of the phase relationship and its stability. The in-phase relationship is
an attractor state that is more stable than any other phase relationship between the tapper and the beat.
There is another phase relationship—anti-phase (i.e. half-way between two consecutive phase positions)—
which is also relatively stable, but (as we will see) its attractiveness is weaker than for the in-phase
behaviour under certain conditions. Automatic corrections of larger asynchronies that are further away
from the in-phase or anti-phase states are more easily suppressed.
When synchronizing to a beat with a uctuating tempo, it is theoretically conceivable that the tapper could C13.P12
correct all asynchronies between the signal and their taps using phase correction alone. However, a more
e ective approach would be to allow for correction of the tapping period as well. In contrast to phase
correction, period correction is not automatic, because it requires the detection of di erences between pairs
of onsets as opposed to individual onsets, and it has been shown to be signi cantly a ected by both
awareness and attention (Repp and Keller 2004). Therefore, period correction is of a second order to phase
correction and thus requires more cognitive resources. Nonetheless, synchronizing to a tempo-changing

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
beat happens quickly and relatively e ortlessly.

Beatmatching in DJ Practice C13.S2

Most percussive musicians, such as pianists and drummers, directly improve the synchrony of their playing C13.P13
with the beat—whether expressed in their own playing (self-entrainment) or in another’s (interpersonal
entrainment)—by correcting the phase and period of movements on the instrument. Entrainment and
sensorimotor synchronization on percussive instruments are thus a direct, albeit signi cantly more
sophisticated, extension of the nger-tapping measured in psychological experiments. Also, non-
p. 239 percussive instrumentalists, such as string players and vocalists, directly adjust their body movements
1
(moving the bow a little earlier, activating their vocal chords a little later) to align with the beat. DJs playing
turntables, however, cannot directly produce the rhythmic patterns of the music they are playing. Rather,
the rhythms and the beat of the music are prerecorded onto vinyl discs, and the DJ has to manipulate their
speed and position on the turntables to get them to synchronize.

Di erent DJs working in di erent dance music scenes will adopt di erent ways of synchronizing, mixing, C13.P14
and stringing records together. In more eclectic scenes, where DJs mix music from several dance music
styles, sometimes with di erent metric structures as well as di erent tempi (e.g. salsa, hip-hop, house, and
funk), the transition from one track to another may be achieved through manipulating the relative loudness
of certain frequency bands (EQing), speaking on a microphone (MCing), or playing an atmospheric part of a
record (with no beat), directing attention towards or away from certain elements of the rhythmic structure
2
to smoothen the change. DJs who strictly play house and/or techno tend to mix in a way that is as
seamlessly synchronized as possible. While many DJs now use digital systems such as the CDJ, which
provides additional visual and numerical feedback and can synchronize recordings automatically via beat-
detection algorithms, mixing records synchronously by ear on analogue turntables is considered a
foundational skill for a DJ. The two main techniques used to synchronize vinyl records with di erent tempi
are slip-cueing and beatmatching. Slip-cueing refers to the process of nding the beat on a track so that it
can be played in time with the beat of the track already playing. Beatmatching is a procedure in which the
3
tempo of the new track is adjusted to match that of the currently playing track. Standard DJ turntables
include a continuous pitch controller that can be used to adjust playback within +8% or –8% of the recorded
speed. Some of these uses were intended by the manufacturers, and some have emerged from DJ practice
4
itself. In entrainment terms, slip-cueing is analogous to phase correction and beatmatching to period
correction.

Most of the work needed to synchronize a new track with the already playing track is done in a way that is C13.P15
only audible to the DJ over headphones, while the audience enjoys dancing to the already playing track. Slip-
cueing is done by pushing the new record back and forth on a slip mat in order to move it into alignment
5
with the track already playing. DJs will often listen out for a particularly noticeable sound in the track. In
House or Techno, this can be the rst of a ‘four on the oor’ bass-drum beat, since it is often the most
dominant of beat markers, while hip-hop DJs are known to slip-cue on the backbeat snare, since it is often
more distinct in a hip-hop track. Some DJs listen to the hi-hats because they are often more audible above
the other instruments in the mix. The DJ might rst listen to a few bars of the track (not necessarily at the
beginning) in order to identify the beat that they intend to slip-cue, and then move the record on the slip
mat so that the needle is situated at the onset of the beat. The DJ then has to either stop the turntable or keep
holding the record in place until they are ready to release it. Before the record is released, it is common to
move the record back and forth with the beat just below the needle, so as to get a feel for where exactly on
the record the beat starts. When the DJ has a feel for where it is, they will release the record—or slip it—at a
6
time point so that it plays in time with the beat of the track playing on the other deck.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
p. 240 Before the DJ makes the new track audible to the audience (let us call it track B), they will want to make sure C13.P16
that it is completely synchronized (with what we will call track A). Slip-cueing is usually not su cient,
because the tempos of the records may be di erent. Even quite small tempo di erences will gradually
accumulate, potentially leading to what is known as a ‘train wreck’—a misalignment of the two tracks that
is so noticeable that it thwarts the enjoyment of the groove and makes dancing to the beat di cult. In order
to prevent this, the DJ has to adjust the speeds of the records so that the tempi of the tracks are aligned. This
beatmatching is achieved by moving the pitch controller. In addition, the DJ will also perform an action
known as ‘nudging’, which involves manually pushing the record gently forward, by pushing either onto the
record itself or on its side. This process can compensate for any phase di erence between the two tracks.
Adjusting the speed and nudging are often done incrementally, in a stepped fashion, but it must be
monitored continuously, especially on analog equipment that can be a ected by random drift in speed.

Di erent DJs will use the pitch controller and nudging in di erent ways to achieve beatmatching. But rst, C13.P17
they have to identify whether the tempo of track B needs to be slowed down or sped up, and this is done by
7
listening to the combined tracks continuously and monitoring the alignment. If record B is slower than
record A, one technique is to turn the pitch control all (or almost all) the way toward you, and then quickly
8
back in the opposite direction—not to its original position, but to a position slightly before. By doing this,
the position of record B is rst moved ahead to catch up with track A and then set to a slightly faster tempo.
The DJ will then listen until the tracks drift into a periodic relationship. If track B is still too slow, the
process is repeated in smaller and smaller intervals (overshooting and undershooting) until a satisfactory
9
level of synchrony is achieved. If phase position is a ected during this process, a slight nudge may help.
The goal is to get the two tracks synchronized before track B is introduced to the audience, although if the DJ
discovers smaller asynchronies after track B is introduced, either of the two tempi can be adjusted, either by
very minimal adjustments on the pitch control, nudging, or brie y holding (slowing) the spindle. However,
since both tracks are now audible, this technique can back re by causing noticeable irregularities to the beat
and thwarting the continuation of the groove. Ideally, all adjustments should be completed before the new
track is introduced, so that the monitoring of beat synchronization can be kept to a minimum and the DJ can
10
focus on other processes, such as blending, ltering (or EQing), and choosing tracks.
Beat Perception and Temporal Correction in DJing C13.S3

From my description of slip-cueing and beatmatching, it should be clear that DJing involves a variety of C13.P18
p. 241 multisensory, cognitive, motor, and musical skills, which can take several years to master. A DJ must be
able to perceive and distinguish between periodicities of simultaneously playing streams of polyrhythmic
sound, monitor levels of synchronicity at the millisecond level, and perform temporal error corrections by
11

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
multi-dextrally manipulating an instrument. Despite the observations that slip-cueing and beatmatching
require a considerable degree of temporal acuity, few studies have investigated the rhythmic abilities of the
12
DJ. Butler and Trainor (2015) compared DJs, percussionists, and non-musicians on a beat mentalizing
task that involved listening to a regular beat pattern followed rst by a period of silence, and then by a
target tone which was either in time or out of time with the previously heard beat. If a target tone was
correctly identi ed as in or out of time, the participant was thought to have accurately maintained the beat
in their mind during the silent period (a process known as ‘beat continuation’). The study found that DJs
were as good as percussionists at correctly identifying target tones, and equally sensitive to the target
tone’s phase o set from the beat. Both DJs and percussionists performed better on the task than non-
musicians. The results suggest that both DJs and percussionists demonstrate sophisticated beat perception
abilities.

In another study, MacCutcheon et al. (2016) tested both beat perception and sensorimotor synchronization C13.P19
capacities in self-trained DJs, formally trained classical string players, and non-musicians. One task
involved simple nger-tapping to a regular metronomic beat, and another task involved nger-tapping to
the metronome while ignoring a simultaneously playing unsynchronized distractor metronome. The
distractor had the same regular tempo but was o set in phase. The results showed that DJs exhibited lower
tapping variability in both tasks, suggesting that they are better at synchronizing to a beat than classically
trained string players and untrained participants. The outperformance of DJs in the distractor task is not
surprising, since attending to and adjusting periodicities with phase asynchronies is crucial to
beatmatching. In addition, the decreased variability during synchronization to a single metronome suggests
that the auditory–motor coordination involved in beatmatching and slip-cueing makes DJs better at
synchronizing than string players more generally.

Slip-cueing, when a record is moved back and forth under the needle and slipped in time to the beat of C13.P20
another record, is perhaps the closest a DJ gets to synchronizing their movements to a beat. DJs of course
often move their bodies and dance to the beat in between or while manipulating the turntables, but while
this may help in perceiving, or feeling, where the beat is (Phillips-Silver and Trainor 2008), these
movements do not actually produce the sounds of the music. The movements involved in slip-cueing are in
some ways similar to moving a bow across the strings of a violin, because the DJ is moving a needle along
the grooves of a record. However, the slipping movement in DJing is an isolated, single movement of lifting
the hand o the record and letting it spin. Of course, the DJ can repeat the slip-cueing if they are unsatis ed
with the records’ alignment, but they have to go through the whole process of identifying the location of the
beat on the record rst.

Furthermore, and unlike for the violinist, any resulting asynchrony cannot be corrected automatically. The C13.P21
p. 242 DJ must either try again or apply some compensatory correction by nudging or holding down the spindle.
In other words, slip-cueing requires voluntary control of phase correction. The DJ cannot rely on the
automaticity that this process allows in normal sensorimotor synchronization, but has to consciously
correct the phase asynchrony. While dancing and moving their whole bodies to the beat may help the DJ feel
where the temporal target is, this does not directly translate to the slip-cueing movement itself.
Importantly, the correction cannot be done instantaneously and repetitively as a continual process on a
beat-by-beat basis, as in simple sensorimotor synchronization. Instead, slip-cueing is one isolated
movement followed by a period of attentive listening to determine if the beats of the two records are
su ciently aligned. If the movement did not align the phase su ciently, the DJ has to start from the
beginning, rewinding the record, and repeat the process until a satisfactory degree of synchrony is achieved.
More conscious control is also required in the period correction of beatmatching. Rather than adjusting the
period directly through corrections of synchronized body movements, the DJ has to manipulate a linear
controller to change the speed of the records. The accumulating asynchrony between two records with
di erent periods must be corrected through repeated and incremental trial and error. Furthermore, the
processes of phase correction (slip-cueing, nudging) and period correction (moving the pitch control) often

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
have to be employed interchangeably, since changes in period necessarily introduces changes in phase. Of
course, for professional DJs with years of practice, these skills become internalized and less e ortful with
time. But they are not automatic in the way that they are in direct sensorimotor synchronization. The
di erence in control of temporal error correction is signi cant, because it suggests that entrainment in DJ
beatmatching operates di erently from other forms of sensorimotor synchronization. In order to
understand this di erence, we must rst understand the role of entrainment in cognitive systems, more
broadly.

Enactive Coupling C13.S4

Entrainment is a special form of coupling that assumes a temporally synchronous relationship between the C13.P22
di erent interacting processes. Both entrainment and coupling more generally are central to explanations
of the embodiment of cognitive systems. There is growing agreement among philosophers of mind and
some cognitive scientists that the mind should not be considered to be purely brain-based. Instead, the
body and the environment in which the body is embedded are thought to take up constitutive roles in the
make-up of mind (Newen et al. 2018). Enactivism is one of the most dominant theories making this claim,
and for enactivists, coupling provides the core mechanism by which cognitive systems become embodied.

The central idea of enactivism, as rst described by Francisco Varela, Evan Thompson, and Eleanor Rosch C13.P23
(1991), is that the mind is not an information-processing machine that is localized in the brain but is an
active process that cannot be localized at all. It is a bidirectional relating of an organism with its
p. 243 environment. Thomas Fuchs likened the enactive relationship between brain, body, and environment to
that of breathing: ‘Just as respiration cannot be restricted to the lungs but only functions in a systemic unity
with the environment, so the individual mind cannot be restricted to the brain’ (Fuchs 2009: 222).
Cognition is thus a dynamic process of reciprocity between a subject and its environment, emerging from
situated sensorimotor experience.

This dynamic reciprocity is engendered by coupling. In physics, coupling simply means the exchange of C13.P24
information between two or more interacting objects or systems. The classic example from mechanics is
two swinging pendulums that are connected by a spring, resulting in the two pendulums’ swinging
behaviour being partly a ected by each other. Enactivists have elaborated the de nition of this process in
ways that highlight the e ects of coupling on interacting systems. According to Di Paolo and De Jaegher
(2012: 4):

Two systems are said to be coupled when parametrical and other structural descriptions of the C13.P25
laws of transformation of states in one of them have a functional dependence on the state variables
of the other, which may be non-linear, piece-wise, state-dependent, and time-varying (in which
case we call the coupling ‘dynamical’). Coupling can be unidirectional or mutual.

Thompson (2007) laid out three criteria for enactive coupling of an agent with the environment: (i) the C13.P26
process is dynamic such that the interacting systems become co-dependent (where changes in the systems
depend mutually on each other); (ii) the mutual interaction results in a structurally coherent suprasystem;
and (iii) that the agent retains its autonomy despite its codependence on the other system. These criteria are
important for positioning coupling as constitutive of cognitive systems, rather than as a result of a causal
relation (from e.g. a central executive, such as the brain).

This coupling is the basis for the constitution of living systems at three levels (Thompson and Varela 2001). C13.P27
At the biological level, the delineation and autonomy of an organism depends on an interconnected network
of environmental and homeostatic processes. This type of organismic organization re ects the metabolic

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
autopoiesis, or self-production, of a living system. Autopoietic and self-organizing systems necessarily give
rise to the most basic sense of identity, a core bodily self (Damasio 1999). Thus, life itself cannot be
separated from mind, and mind is partly de ned by the homeodynamic regulation between a biological
being and its biological context. At the personal level, the mind constitutes a cyclicity between organism and
environment a orded by sensorimotor coupling. Perception is the active engagement of an agent within its
world. To perceive is a function of situated movement and to move is a function of what is perceived: ‘to
know a thing is to have learned how to deal with it’ (Fuchs 2009: 224). The sensorimotor action–perception
loops create the sense of agency—a feeling of the active and goal-oriented self. The nal level of cyclicity in
the mind is the intersubjective. The sensorimotor coupling that occurs between di erent self-organizing
and self-sustaining organisms in social interaction is the basis for how an agent perceives itself as an
13
intersubjective self.

p. 244 Cognition and perception are primarily de ned at the level of sensorimotor cycles in the enactive approach. C13.P28
The fundamentally relational aspect of sensorimotor coupling means that all behaviours and the cognitive
mechanisms that accompany them are highly context-sensitive and thus can never be fully simulated using
algorithms. Cognition happens not because of a series of computations in the brain, but as a set of speci c
active engagements with speci c environments. Perceptual experience is the ‘mastery of lawful
sensorimotor relations proper of the mode of engagement with the environment currently enacted by an
agent’ (Di Paolo et al. 2017: 7). Thus, to perceive an object or an event (e.g. a co ee cup, or a melody) is to
master the movements involved in engaging with that object through the appropriate sensory modalities
(vision, audition, motor, touch, smell, and proprioception).

To explain how cognitive processes and sensorimotor behaviours emerge from the distributed world– C13.P29
body–brain system, many enactivists rely on the framework of systems dynamics (Di Paolo et al. 2017;
Thompson and Varela 2001). Systems dynamics in turn relies on principles from synergetics, which models
the self-organizing behaviour of functional groupings of structural elements, or synergies, as changes in
coordination dynamics (Kelso 2009). According to these theories, biological organisms such as humans are
open non-equilibrium systems sharing important mechanistic properties with natural, physical, and
chemical systems, such as uids and res, where patterns of behaviour emerge due to the coordination
dynamics of the system rather than a central homunculus. Humans are thus seen as self-organized dynamic
systems whose behavioural patterns form as a result of coupled subsystems. The three main forms of
coordination in such systems are absolute coordination, relative coordination, and no coordination. In
oscillatory systems, the coordination amounts to di erent degrees of synchrony between elements, from
phase-locked synchrony to a mix of synchrony and asynchrony, to complete asynchrony.

Armin Fuchs and J. A. Scott Kelso observed that during certain patterns of coordinated behaviour, humans C13.P30
exhibit dynamic properties that are similar to those occurring in other nonlinear self-organizing systems
(Kelso et al. 1987; Schoner and Kelso 1988). Their rst experiments tested the dynamics of wagging a nger
in anti-phase to a regular beat. At lower speeds, humans are relatively good at entraining their ngers in
anti-phase. Recall that the anti-phase position is the second most stable phase position during rhythmic
entrainment. However, if the rate of the isochronous rhythm starts to increase, the synchronized behaviour
becomes increasingly attracted to the in-phase state (consistent with the automaticity of phase correction
and the greater stability of the in-phase position) and, at a critical tempo, will spontaneously switch to
14
moving in phase. Using a model of coupled nonlinear oscillators and stochastic dynamics, Fuchs and Kelso
showed that the coupled dynamic system of the nger-wagging and the speeding isochronous beat starts to
exhibit a particular type of enhanced variability that is also observed in other self-organizing systems.
These properties were later generalized to a greater variety of body movements between di erent limbs,
di erent humans, and di erent animal species (Fuchs and Kelso 2018). In later experiments using
p. 245 electroencephalography and magnetoencephalography, they found that this switching from in-phase to
anti-phase at the behavioural level was mirrored in a switching in the phase relationships of certain
oscillatory components of the brain signal (Kelso et al. 2013). Similar dynamic properties are found when

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
measuring sensorimotor and brain responses of two mutually interacting subjects, suggesting
15
intersubjective coupling in and between behavioural and neural domains. The discoveries revealed the
fundamentally self-organized process of nonlinear reciprocal causalities emerging during spontaneous
switching between anti-phase and in-phase behaviour, and suggests that the process of entraining to a beat
is a result of a complex, self-organized system than emerges from the dynamic coordination of neural,
sensorimotor, and environmental processes.

The Distribution of Entrainment in Rhythmic Practice C13.S5

On the enactivist view, coupling and entrainment are thus the processes that enable the embodied C13.P31
distribution of a cognitive system. Adopting a theory closely related to enactivism, Krueger uses extended
mind theory to argue how entrainment underlies the extension of emotions in music (Krueger 2014). He
explains that when we play, listen, or dance to music, we o oad some of the emotional labour of our own
a ective-regulatory processes onto its rhythmic, harmonic, melodic, and dynamic structures. By
coordinating our attention, movements, intentions, and beliefs through rhythmic entrainment, our
individual or collective emotions can be synchronously organized as well (Krueger 2015). I have suggested
elsewhere that in a dance club, entrainment plays a role in enabling socio-a ective extensions across the
dance oor and the distribution of ‘vibe’ between the music, the DJ, the dancers, and the various other
properties of the club environment (Witek 2019). But rhythmic entrainment is not just a tool by which other
a ective and cognitive processes can be distributed. The mechanisms of entrainment are themselves
distributed, in di erent ways depending on the activity in question.

In motor entrainment, where an agent is synchronizing with an external beat, as in most nger-tapping C13.P32
experiments or when a drummer plays along to a click-track, the information exchange is one-directional,
since only the behaviour of the tapper or drummer is a ected by the music. Nonetheless, the machine
producing the click (e.g. a computer or a mechanical metronome) and the agent form a coherent, distributed
system where both contribute to embodied mind. When a drummer is playing solo and self-entraining to
their own beats, the distribution has a more limited reach, stopping at the body of the drummer (which
16
includes the brain), and the musical instrument. There is only one oscillator, and it entrains to itself. In
both motor and self-entrainment, the distribution happens via the sound waves travelling to the drummer
from the music, click, or drum. In classical mechanics terms, the sound waves are the force, or spring, that
p. 246 connects them. Phase correction happens either automatically (if the phase di erence between the
external beat and the agent’s behaviour is small enough) or through the conscious perception of and motor
adaptation to the perceived asynchronies. If there is a tempo change in the music, the agent adapts by
consciously correcting the period of their coordinated movements.

In social entrainment, such as when two drummers play together, the system is also distributed via C13.P33
soundwaves, but now the information exchange and dependence is fully mutual. Each drummer a ects and
is a ected by the other, and the cognitive system of their ensemble emerges from the bidirectional
coordination of their sensorimotor systems and their dynamic attending processes. The distribution covers
the bodies, brains, and instruments of more than one agent. Experiments have shown that when two
tappers are asked to synchronize and keep a steady beat together, their coupled system is continuously and
mutually adapting in such a way that their temporal corrections oscillate in opposite directions on a tap-by-
tap basis (Konvalinka et al. 2010). In other words, the phase and period corrections are distributed across
the two agents in the dyad.

When beatmatching, the DJ is interacting with analog turntables, which are a kind of electric motor. The C13.P34
oscillations of the records are a result of electrical energy converted into mechanical energy. In other words,
they are not self-sustained oscillators in the way that living systems are, as in the social entrainment

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
scenario described above. They are more like the click track, but only if the click is produced by a physical
object whose timing can be manipulated (e.g. a mechanical metronome). In order to synchronize the two
records, the DJ has to rst entrain their attention to one of them. It is unlikely that the DJ can entrain to both
at the same time, unless their periods are already matched and they are in a simple phase ratio relationship
17
to one another. However, the DJ can switch their entrained attention between them. Most likely, they
entrain mostly to the record that is already playing (record A) and correct the record that is being mixed in
(record B). This attentional entrainment gives the DJ the reference framework to which the beat of record B
must be matched. However, while the attentional phase correction to record A may be automatic, it requires
constant attention to sustain. An important skill in DJing is to be able to attend to and perceive the timing of
another record that has di erent phase and period without falling out of the coupling with the master
record (record A). The more experience the DJ has, the better they are able to simultaneously attend to more
than one periodicity while staying entrained to one.

But there is also another form of coupling happening in the beatmatching DJ system: the coupling between C13.P35
the two records. As we have seen, to synchronize record B to record A, the DJ adjusts the period using the
pitch control and the phase by slip-cueing and nudging. One might therefore see the DJ as enacting the
driving force that entrains the two records. The DJ is the spring that connects the two oscillating records,
and the information exchange happens via their skilful sensorimotor engagement with the turntables. The
behaviour of record B depends on the timing of record A, but only because the DJ physically enacts the phase
and period correction of entrainment. The DJ must actively drive the synchronization of two external
independent oscillators. Thus, beatmatching in DJing introduces a new form of entrainment that is di erent
p. 247 from those already described in the literature (Phillips-Silver et al. 2010); it is a mixture of physical (the
turntables), attentional, neural, and sensorimotor processes (the DJ). Without the DJ operationalizing
temporal correction in this way, the records would not synchronize. It also reveals the embodied and
environmental distributedness of entrainment in a new way. The phase and period corrections happen in the
world, in the interaction of the DJ with the turntables. If we stand close enough to the DJ, we can literally see
the corrections happening (although unless we wear headphones that are connected to the mixer, or the DJ
plays both tracks to the audience over speakers, we cannot hear them). In normal sensorimotor
synchronization, the corrections are also happening in the world, but we only see and hear the temporal
result of the corrections. In DJing, beatmatching physically enacts and externalizes the temporal correction
mechanisms that are normally ‘invisible’ in sensorimotor synchronization. In this way, the beatmatching
DJ is an embodied model of rhythmic entrainment.
Conclusion C13.S6

Thus, what DJ beatmatching shows us is not that entrainment in this musical practice is more embodied or C13.P36
distributed than in any other musical or non-musical behaviour, such as playing the drums, marching in a
parade, or taking part in a spoken conversation. And since, according to enactivism, the mind itself emerges
from coupling at biological, sensorimotor, and social levels, our minds and lives are always already

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
distributed. However, in most situations, we don’t see or hear the distribution happening, only the result of
the information exchange. This normally ‘hidden’ nature of the distribution of coupling may explain why
the mind and its information processing is so often thought to be con ned to the insides of the brain, and
why dualistic models of the mind can seem intuitive.

Instead, what this analysis of DJ beatmatching shows is an unusually vivid form of distributed entrainment C13.P37
in which we can both hear and see not just the results of the temporal corrections, but the actual corrections
themselves. It shows us more explicitly the distributed nature of the human mind and behaviour. Adjusting
the pitch controller and nudging the spinning records are physical enactions of the same kinds of temporal
corrections that happen when we tap our feet to a beat, when swinging pendulums entrain, and when ring
patterns phase-lock to each other in neuronal ensembles. In this way, the mind is a DJ all the time, with its
cognitive mechanisms always actively distributed across the brain, body, and the environment.

Notes
1. Of course, sometimes musicians purposefully resist perfect synchrony, as when playing with microtemporal C13.N1
sophistication (Iyer 2002).

p. 248 2. These techniques could also be considered as contributing to entrainment—albeit a looser, more malleable kind. C13.N2

3. Beatmatching is also sometimes referred to as ʻbeatmixingʼ or ʻbeatsynchingʼ. As DJing developed into a global music C13.N3
practice, di erent local scenes have adapted their own terminology for the techniques, skills, and practices involved.

4. Scratching—a hip-hop practice where a record is manually pushed back and forth on a slip mat to create new rhythmic C13.N4
and melodic pattern—is perhaps the most virtuosic skill resulting from the adaptation of record players for DJ practice.

5. While normal rubber vinyl mats ensure that the record moves in synchrony with the rotating decks, a coarse felt slip mat C13.N5
allows the record to slide on top of the mat, such that the record be can stopped and moved while the platter of the direct
drive turntable rotates underneath.

6. This whole process is still only heard by the DJ through headphones. C13.N6

7. The DJ may also know the tempo of the track from previous listening. C13.N7

8. On the pitch controller, pushing the controller toward you makes the record spin faster, while pushing away makes it C13.N8
slower.

9. Phase misalignment can be perceived not only as a time di erence but also as a timbral quality. Playing two identical (or C13.N9
near-identical) sounds, such as the hi-hats of two di erent records, simultaneously but with a small (i.e. less than 20 ms)
phase di erence, will cause flanging, which sounds like a sweeping filter e ect. Some DJs will use this e ect as a sign of
misalignment between the two signals.

10. Blending (or mixing) is the process of increasing the volume of the new tracks so it is audible to the audience on top of the C13.N10
other track, making a smooth transition between the two. Filtering or EQing involves using controllers on the turntables to
increase or decrease the volume of certain frequency bands of the signal. These techniques can be as important during a
transition as getting the records synchronized through beatmatching. E.g. if the melody or vocals of the two records clash
harmonically, blending and filtering can help smoothen the dissonant transition between them (although DJs sometimes
seek out such dissonances in their mixing).
11. Nonetheless, there is a history of disputing the musical sophistication of DJs, based on the common misconception that C13.N11
all they do is ʻplay tracksʼ. While the skills of a DJ may indeed be di erent from those of other instrumentalists, they are to
the last degree musical (see MacCutcheon et al. 2016 for a qualitative study of the perception of DJ musicality).

12. See Greasley and Prior (2013) for a study on DJs perception of shape. C13.N12

13. The three levels of enactive cycles are not to be thought of as hierarchically organized in a way that one (e.g. the C13.N13
biological) is more fundamental than another (e.g. the social): An organismʼs sense of self is an inextricable mixture of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
metabolic, sensorimotor, and intersubjective relationships with the environment and other subjects in it.

14. When humans start o wagging in phase, there is no equivalent switch to anti-phase, revealing asymmetry in the stability C13.N14
of the two states.

15. At the critical point, the scaling of both behavioural and neural variability fluctuations follows a common law of 1/f. C13.N15

16. However, this one oscillator could be broken down into several metrically related sub-oscillators, depending on the metric C13.N16
and rhythmic structure of the music being played.

17. It may be that a single entrained attentional oscillator with a lower amplitude and wider attentional focus could C13.N17
p. 249 encompass two minimally unsynchronized tracks (Danielsen 2010; 2018); but because this looser entrainment results
in less precise phase expectations, it is unlikely that it would enable accurate correction of the asynchrony between the
tracks.
References C13.S7

Barnes, R., and Jones, M. R. (2000). Expectancy, attention and time. Cognitive Psychology 41(3): 254–311. C13.P38
Google Scholar WorldCat

Butler, B. E., and Trainor, L. J. (2015). The musician redefined: A behavioural assessment of rhythm perception in professional C13.P39
club DJs. Timing and Time Perception 3(1–2): 116–132.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Clayton, M., Sager, R., and Will, U. (2004). In time with the music: The concept of entrainment and its significance for C13.P40
ethnomusicology. European Meetings in Ethnomusicology 1: 1–86.
Google Scholar WorldCat

Cook, P., Rouse, A., Wilson, M., and Reichmuth, C. (2013). A California sea lion (Zalophus californianus) can keep the beat: Motor C13.P41
entrainment to rhythmic auditory stimuli in a non-vocal mimic. Journal of Comparative Psychology 4: 412–427.
Google Scholar WorldCat

Damasio, A. (1999). The feeling of what happens: Body, emotion and the making of consciousness. Vintage. C13.P42
Google Scholar Google Preview WorldCat COPAC

Danielsen, A. (2010). Here, there and everywhere. Three accounts of pulse in DʼAngeloʼs ʻLe and Rightʼ. In A. Danielsen (ed.), C13.P43
Musical rhythm in the age of digital reproduction, 19–35. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Danielsen, A. (2018). Analysing beat bin metre in neo soul grooves. In C. Scotto, K. Smith, and J. Brackett (eds), The Routledge C13.P44
companion to popular music analysis: expanding approaches, 179–189. Routledge.
Google Scholar Google Preview WorldCat COPAC

Di Paolo, E., Buhrmann, T., and Barandiaran, X. (2017). Sensorimotor life: An enactive proposal. Oxford University Press. C13.P45
Google Scholar Google Preview WorldCat COPAC

Di Paolo, E., and De Jaegher, H. (2012). The interactive brain hypothesis. Frontiers in Human Neuroscience 6: 163. C13.P46
https://doi.org/10.3389/fnhum.2012.00163
Google Scholar WorldCat

Fuchs, A., and Kelso, J. A. S. (2018). Coordination dynamics and synergetics: From finger movements to brain patterns and ballet C13.P47
dancing. In S. Muller, P. Plath, G. Radons, and A. Fuchs (eds), Complexity and synergetics, 301–316. Springer.
Google Scholar Google Preview WorldCat COPAC

Fuchs, T. (2009). Embodied cognitive neuroscience and its consequences for psychiatry. Poiesis und Praxis 6(3–4): 219–233. C13.P48
Google Scholar WorldCat

Gámez, J., Yc, K., Ayala, Y. A., Dotov, D., Prado, L., and Merchant, H. (2018). Predictive rhythmic tapping to isochronous and tempo C13.P49
changing metronomes in the nonhuman primate. Annals of the New York Academy of Sciences 1423: 396–414.
Google Scholar WorldCat

Grahn, J. A. (2012). Neural mechanisms of rhythm perception: Current findings and future perspectives. Topics in Cognitive C13.P50
Science 4(4): 585–606.
Google Scholar WorldCat

Greasley, A. E., and Prior, H. M. (2013). Mixtapes and turntablism: DJsʼ perspectives on musical shape. Empirical Musicology C13.P51
Review 8(1): 23–43.
Google Scholar WorldCat
Honing, H. (2012). Without it no music: Beat induction as a fundamental musical trait. Annals of the New York Academy of C13.P52
Sciences 1252(1): 85–91.
Google Scholar WorldCat

Iyer, V. (2002). Embodied mind, situated cognition, and expressive microtiming in African-American music. Music Perception C13.P53
19(3): 387–414.
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
Jones, M. R. (2009). Musical time. In S. Hallam, I. Cross, and M. Thaut (eds), The Oxford handbook of music psychology, 81–92. C13.P54
Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Jones, M. R., and Boltz, M. (1989). Dynamic attending and responses to time. Psychological Review 96(3): 459–491. C13.P55
Google Scholar WorldCat

p. 250 Kelso, J. A. S. (2009). Coordination dynamics. In R. A. Meyers (ed.), Encyclopedia of complexity and systems science, 1537–1564. C13.P56
Springer.
Google Scholar Google Preview WorldCat COPAC

Kelso, J. A. S., Dumas, G., and Tognoli, E. (2013). Outline of a general theory of behaviour and brain coordination. Neural C13.P57
Networks 37: 120–131.
Google Scholar WorldCat

Kelso, J. A. S., Schöner, Z., Barnea, Z., and Haken, H. (1987). Phase-locked modes, phase transitions and component oscillators in C13.P58
biological motion. Physica Scripta 35(1): 79–87.
Google Scholar WorldCat

Konvalinka, I., Vuust, P., Roepstor , A., and Frith, C. D. (2010). Follow you, follow me: continuous mutual prediction and C13.P59
adaptation in joint tapping. Quarterly Journal of Experimental Psychology 63(11): 2220–2230.
Google Scholar WorldCat

Krueger, J. (2014). A ordances and the musically extended mind. Frontiers in Psychology 4: 1003. C13.P60
https://doi.org/10.3389/fpsyg.2013.01003
Google Scholar WorldCat

Krueger, J. (2015). The a ective ʻweʼ: Self-regulation and shared emotions. In T. Szanto and D. Moran (eds), The phenomenology C13.P61
of sociality, 263–277. Routledge.
Google Scholar Google Preview WorldCat COPAC

Large, E. W., and Jones, M. R. (1999). The dynamics of attending: How people track time-varying events. Psychological Review C13.P62
106(1): 119–159.
Google Scholar WorldCat

Large, E. W., and Kolen, J. (1994). Resonance and the perception of musical meter. Connection Science 6(1): 177–208. C13.P63
Google Scholar WorldCat

MacCutcheon, D., Greasley, A. E., and Elliott, M. T. (2016). Investigating the value of DJ performance for contemporary music C13.P64
education and sensorimotor synchronisation (SMS) abilities. Dancecult 8(1): 46–72.
Google Scholar WorldCat

Newen, A., De Bruin, L., and Gallagher, S. (eds) (2018). The Oxford handbook of 4E cognition. Oxford University Press. C13.P65
Google Scholar Google Preview WorldCat COPAC

Patel, A. D., Iversen, J. R., Bregman, M. R., and Schulz, I. (2009). Experimental evidence for synchronization to a musical beat in a C13.P66
nonhuman animal. Current Biology 19(10): 827–830.
Google Scholar WorldCat

Phillips-Silver, J., Aktipis, C. A., and Bryant, G. A. (2010). The ecology of entrainment: Foundations of coordinated rhythmic C13.P67
movement. Music Perception 28(1): 3–14.
Google Scholar WorldCat

Phillips-Silver, J., and Trainor, L. J. (2008). Vestibular influence on auditory metrical interpretation. Brain and Cognition 67(1): C13.P68
94–102.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469884 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Repp, B. H. (2002a). Automaticity and voluntary control of phase correction following event onset shi s in sensorimotor C13.P69
synchronization. Journal of Experimental Psychology: Human Perception and Performance 28(2): 410–430.
Google Scholar WorldCat

Repp, B. H. (2002b). Phase correction in sensorimotor synchronization: Nonlinearities in voluntary and involuntary responses to C13.P70
perturbations. Human Movement Science 21(1): 1–37.
Google Scholar WorldCat

Repp, B. H., and Keller, P. E. (2004). Adaptation to tempo changes in sensorimotor synchronization: E ects of intention, attention, C13.P71
and awareness. Quarterly Journal of Experimental Psychology Section A: Human Experimental Psychology 57(3): 499–521.
Google Scholar WorldCat

Repp, B. H., and Su, Y.-H. (2013). Sensorimotor synchronization: A review of recent research (2006–2012). Psychonomic Bulletin C13.P72
and Review 20(3): 403–452.
Google Scholar WorldCat

Schoner, G., and Kelso, J. A. S. (1988). Dynamic pattern generation in behavioural and neural systems. Science 239(4847): 1513– C13.P73
1520.
Google Scholar WorldCat

Thompson, E. (2007). Mind in life: Biology, phenomenology and the sciences of mind. Harvard University Press. C13.P74
Google Scholar Google Preview WorldCat COPAC

Thompson, E., and Varela, F. J. (2001). Radical embodiment: Neural dynamics and consciousness. Trends in Cognitive Sciences C13.P75
5(10): 418–425.
Google Scholar WorldCat

Toiviainen, P., Luck, G., and Thompson, M. (2009). Embodied metre: Hierarchical eigenmodes in spontaneous movement to C13.P76
music. Music Perception 28(1): 59–70.
Google Scholar WorldCat

p. 251 Varela, F. J., Thompson, E., and Rosch, E. (1991). The embodied mind: Cognitive science and human experience. MIT Press. C13.P77
Google Scholar Google Preview WorldCat COPAC

Wing, A. M., and Kristo erson, A. B. (1973). Response delays and the timing of discrete motor responses. Perception and C13.P78
Psychophysics 14(1): 5–12.
Google Scholar WorldCat

Witek, M. A. G. (2019). Feeling at one: Socio-a ective distribution, vibe and dance music consciousness. In D. Clarke, R. Herbert, C13.P79
and E. F. Clarke (eds), Music and consciousness 2, 93–112. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Witek, M. A. G., Popescu, T., Clarke, E. F., Hansen, M., Konvalinka, I., Kringelbach, M. L., and Vuust, P. (2017). Syncopation a ects C13.P80
p. 252 free body-movement in musical groove. Experimental Brain Research 235(4): 995–1005.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
CHAPTER

14 Non-isochronous Metre in Music from Mali  C14

Rainer Polak

https://doi.org/10.1093/oxfordhb/9780190947279.013.22 Pages 253–C14.P117


Published: 08 December 2021

Abstract
The basic building blocks for rhythmic structure in music are widely believed to be universally con ned
to small-integer ratios. In particular, basic metric processes such as pulse perception are assumed to
depend on the recognition and anticipation of even, categorically equivalent durations or inter-onset
intervals, which are related by the ratio of 1:1 (isochrony). Correspondingly, uneven (non-isochronous)
beat subdivisions are theorized as instances of expressive microtiming variation, i.e. as performance
deviations from some underlying, categorically isochronous temporal structure. By contrast,
ethnographic experience suggests that the periodic patterns of uneven beat subdivision timing in
various styles of music from Mali themselves constitute rhythmic and metric structures. The present
chapter elaborates this hypothesis and surveys a series of empirical research projects that have found
evidence for it. These ndings have implications for metric theory as well as for our broader
understanding of how human perception relates to cultural environments. They suggest that the bias
towards isochrony, which according to many accounts of rhythm and metre underlies pulse
perception, is culturally speci c rather than universal. Claims regarding cultural diversity in the study
of music typically concern styles and meanings of performance practices. In this chapter, I will claim
that basic structures of perception can vary across cultural groups too.

Keywords: categorical rhythm perception, metric pulse, beat subdivision, isochrony, expressive timing,
enculturation, Mali
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

THE present chapter concerns the uneven subdivision of metric beats in various traditional forms of music C14.P1
from Mali. In this repertoire, beats are strictly isochronous, but subdivisions often seem to be governed by
ratios that fall somewhere in between isochrony (1:1) and a shu e rhythm (2:1). From the commonly
accepted point of view which equates metric regularity with isochrony, such uneven beat subdivisions are
generally regarded as expressive performance deviations from some underlying, structurally isochronous
reference framework. Indeed, this concept of expressive microtiming variation or participatory discrepancy
(Keil 1987) represents the standard interpretation of the well-researched timing of the ‘swing eighths’ in
jazz performance (Benadon 2006; Butter eld 2011; Honing and de Haas 2008; Prögler 1995; Wesolowski
2015). By contrast, my long-term, participatory ethnographic experience with jembe-centred, dance-
oriented drum ensemble music from southern Mali has convinced me that, in the context of this musical
culture, uneven subdivisions, which are governed by ratios other than 1:1, may constitute temporal
references structures in their own right. This chapter will (1) consider the roles that isochrony and non-
isochrony play in various theories of rhythm and metre, (2) elaborate the hypothesis of non-isochronous

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
subdivision-based metre, and (3) provide summaries of ve empirical research projects that incrementally
provide evidence for the hypothesis. Finally (4), I will summarize the empirical ndings and discuss some of
their implications for the psychology of rhythm perception, suggesting a strong role of culture and the
speci cs of cultural environments for perceptual capacities.

p. 254
Isochrony and Non-isochrony in Theories of Rhythm and Metre C14.S1

Most theories of rhythm and metre focus on the relative proportions between durations or inter-onset C14.P2
intervals. Absolute time values certainly also play strong roles in such theories, as the production and
perception of rhythmic gures and metric types depend on event density and tempo (Johansson 2009;
London 2012). However, the structural constituents of these gurations are typically conceived of, and
symbolically represented, as temporal ratios. The speci c ratios constitutive of human rhythm and metre
perception are widely believed to be quite constrained in terms of their mathematical complexity. Many
theories contend that rhythmic categories and their metric relationships are tied to the smallest integer
ratios, 1:1 and 2:1. The former (1:1) is particularly relevant for theories of metre, insofar as a human
tendency to recognize and anticipate successions of categorically isochronous durations is assumed to
1
underpin basic metric processes, such as pulse perception. This assumption that metric beat requires
isochrony (i.e. successive 1:1 ratios) is not con ned to theories of metre in Euro-American music (e.g.
Lerdahl and Jackendo 1983) but is also prominent in comparative (Hood 1971; Savage et al. 2015) and
Africanist ethnomusicology (Agawu 2006; Arom 1984; Burns 2010; Kubik 1988; Locke 1982; 2010; Nketia
1974; Waterman 1952). The psychological theory of dynamic attending (Jones and Boltz 1989; Large and
Jones 1999) suggests that humans tend to entrain to isochronous periodicities which they perceive as
simple and as underlying other, more complex rhythms in their environment. Biomusicologists have argued
that this tendency represents an evolved predisposition that is particularly advantageous when individual
participants in social gatherings synchronize their behaviours by entraining to the same beat, as in the case
of music for ritual and dance (Fitch 2012; 2013; Merker et al. 2009; Ravignani et al. 2016; Ravignani and
Madison 2017). The assumption of a biological predisposition for isochrony has been reinforced by advances
in neuroscience, speci cally the proposition that isochronous beat and metre perception may be directly
controlled by mechanisms of resonance with neural oscillations (Large 2008; Large and Kolen 1994; Large
and Snyder 2009; Tal et al. 2017). In summary, the perception of isochrony is widely credited as the
fundamental condition of metric beat perception.

The psychological theory of categorical rhythm perception further bolsters these views. According to C14.P3
categorical perception, our minds structure in nitely nuanced, continuous sensory spaces (e.g. light, pitch,
speech) by grouping segments of the respective continua into a speci c number of discrete, quasi-symbolic
categories (e.g. colours, tones, phonemes). The di erence between percepts within categories is diminished
(compression), whereas it is increased at the border between two categories (separation): The same physical
di erence, such as the range between two sonic frequencies, appears larger when two percepts fall into two
distinct categories than when they fall into the same one; that is, the mechanism which achieves the
p. 255 categorical structuring of perception consists of a warping of the perceptual space. The key function of
categorical perception is to reliably and uently identify things and events and thus endow organisms with
an e cient perceptual grasp on their in nitely complex environments (Goldstone and Hendrickson 2010;
Harnad 1990; Repp 1984). Music psychological research, through experimental testing of the identi cation
and reproduction of simple rhythms, has suggested that rhythm perception basically works with only two
fundamental categories, even and uneven, which are characterized by prototypes near the two smallest
integer ratios, 1:1 and 2:1 (Clarke 1987; Fraisse 1956; 1982; Povel 1981). More complex rhythmic structures
are assumed to be constructed from these basic categories, comparable to the way Western musical sta
notation is able to represent the majority of complex rhythms with symbols that basically express
categorical 1:1 and 2:1 relations (additionally using diacritical marks for the apparently less fundamental

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
ratio of 3:1).

An important component of the theory of categorical rhythm perception is the distinction between C14.P4
categorical structure and performance timing. Human performers produce both intentional and
unintentional timing and tempo uctuations. These so-called microtiming or microrhythmic variations are
assumed to be perceived as the deviation of a performed rhythm from the perceptual prototype of a
category, and these divergences are theorized as making rhythm more expressive, interesting, and
pleasurable (Clarke 1985; 1987; 1999; 2000; Clarke and Do man 2014; Keil 1987). A perceptual prototype is
the focal representation of a category which biases the perception of all category members—a phenomenon
that has been addressed as perceptual attractor (Repp et al. 2012) or magnetic e ect (Feldman et al. 2009).
This perceptual magnet or attractor e ect has been robustly demonstrated in sensorimotor synchronization
(‘tapping’) studies. The participants’ task in this experimental paradigm is to tap along with simple
rhythms in best possible synchrony. The di erence between the target or pacing rhythm and response,
measured either during the synchronization phase or during a continuation phase that follows after the
rhythm stops being played in the participant’s headphone, allows researchers to evaluate the participants’
mental representation of the target rhythms. Undistorted synchronization/continuation of the target
rhythm indicates that the given ratio ts the prototype of a rhythmic category in the participant’s
perception. By contrast, consistent distortion occurs when the pacing rhythm does not directly t any of the
participant’s perceptual prototypes but is a ected by the attraction of a neighbouring prototype.

Tapping studies found a strong bias towards a prototypical 2:1 ratio for all uneven two-element rhythms C14.P5
(see Figure 14.1). Higher ratios, such as 3:1, were consistently lowered towards 2:1—i.e. the relatively sharp
rhythmic contrast between the two elements was reduced. By contrast, less uneven ratios such as 3:2 were
enlarged in the direction of 2:1, involving an increase of the rhythmic contrast. Only the attest end of
uneven rhythms was further straightened out towards 1:1 (see e.g. Essens and Povel 1985; Jacoby and
McDermott 2017; Povel 1981; Repp et al. 2005; 2011). The participant groups used in these studies included
highly trained musicians; the perceptual bias towards 2:1 thus does not seem to depend on degrees of
p. 256 musical skill and expertise. However, all participants were from Western Europe or North America,
suggesting that a Western cultural bias might play a role.
Figure 14.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Upper part: Selection of small-integer ratios illustrating various two-element rhythms, given in four di erent formats (from le to C14.F1
right): small-integer ratio, percentages of the period, quotient of the long-short ratio, and quotient of short-long ratio. Lower
part: Schematic representation of the perceptual categorization of the space of two-element rhythms according to the current
state of research. Arrows roughly summarize the main biases found in classical tapping studies using Western listeners (e.g.
Povel 1981; Essens and Povel 1985; Repp et al. 2005, 2011). The histogram shows the distribution of participantsʼ stabilized
responses in a recently developed iterative tapping paradigm (taken from Jacoby and McDermott 2017, Figure S1). Horizontal
brackets schematically indicate categories suggested by this research, with prototypes at 1:1, near 2:1, and near 3:1.

Adapted from Polak et al., 2018, Figure 1.

In conclusion, an impressively broad cluster of theories support the conclusion that the smallest integer C14.P6
ratios, 1:1 and 2:1, are of fundamental importance as the perceptual primitives from which more complex
rhythmic and metric structures are constructed.

However, theories of rhythm and metre do not form a monolithic block in this respect. For instance, Justin C14.P7
London’s metric theory allows for a degree of non-isochrony in metric pulse-trains, and proposes that this
degree of non-isochrony is limited by the mathematical principle of maximal evenness, which requires that
a given number of beats must be distributed in a given span of a subdivision cycle in the most evenly
distributed way that is available (London 2012). Thus, three metric beats in the context of a cycle spanning
seven subdivisions need to be distributed as 3:2:2 (or one of its rotations, 2:2:3 or 2:3:2) to be perceivable as
a metric pulse-train; by contrast, 4:2:1 or 3:3:1 would violate this constraint on metric well-formedness.
Maximally even non-isochronous beat cycles such as 2:2:3 do exist and function as main metric reference
level, for instance in dance music from the Balkans (Goldberg 2017). While metric types of this kind typically
p. 257 appear odd, irregular, and di cult from the perspective of most Western European and North American
theorists, musicians, and listeners alike, cross-cultural experiments have demonstrated that enculturated
listeners familiar with music from the Balkans, North Africa, and Turkey, among others places, do not
necessarily nd them more di cult to perceive than isochronous metres (Drake and El Heni 2003; Hannon
et al. 2012; Kalender et al. 2013; Yates et al. 2017). Importantly, Western infants at the age of six months are
equally well disposed to perceive isochronous and non-isochronous metres. This nding suggests that the
bias towards isochronous metres shown by Euro-American 12-month-old infants, children, and adults,
which increases in stability with age, is a result of their speci c enculturation in a musical environment that
privileges isochronous metre, and is not a biological disposition (Hannon and Trehub 2005a; 2005b).
Studies that thus productively integrate cross-cultural and developmental perspectives are highly relevant
insofar as they directly contradict the near-axiomatic assumption that the widespread usage of 1:1 and 2:1 as
basic building blocks for temporal reference structures arises from a universal constraint limiting the usage
of more complex ratios. Perhaps the most radical questioning of theories that assume categorical isochrony
underlying pulse and metre originates from studies in Scandinavian folk-dance music, which demonstrate
the simultaneous co-occurrence of non-isochronous timing patterns at both the beat and the subdivision
levels (Johansson 2017; Kvifte 2007). This casts doubts on the argument that an isochronous fastest pulse is
cognitively necessary in order to provide a common denominator to which all events in a rhythmic surface
can be mapped. Alternatively known as ‘density referent’, ‘fastest pulse’, or ‘elementary pulse’ in
ethnomusicology (Hood 1971; Koetting 1970; Kubik 1988; Nketia 1974) and ‘metric oor’ in music theory
(London 2012), this concept is required for London’s explanation of the well-formedness of non-
isochronous beat-cycles by the principle of maximal evenness, which assumes a common greatest divisor

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
for coordinating the maximally even distribution of uneven beats. However, no such isochronous common
greatest divisor is available for metric perception when the metric oor itself is structurally non-
isochronous. Non-isochronous subdivision thus represents a critical test case for the perceptual possibility
of non-isochronous metre. Its existence would challenge not only Euro-American music-theoretical
concepts, and not only theories of isochronous metre, but also ethnomusicological theories of African
rhythm, and current pulse-based theories of non-isochronous metre. The jembe ensemble music of Mali
provides many examples of non-isochronous subdivision patterns, and this music provides the empirical
basis for the hypothesis being proposed here.

The Hypothesis of Non-isochronous Subdivision-Based Metre C14.S2

Non-isochronous beat subdivisions occur in a broad array of genres and styles from Mali and neighbouring C14.P8
p. 258 countries. For example, Video Sample 1 features a trio recording of jembe-centred drum ensemble music.
(The video is accessible on the handbook’s accompanying website .) The ensemble features two goblet-
shaped jembe drums, played with bare hands, and one dundun, a cylindrical drum, beaten with a stick. The
rst jembe serves as the lead drum, which controls the intensity and extemporizes the musical form of the
piece in performance; the second jembe’s accompaniment provides time-keeping, metric density, and feel
by a simple ostinato; and the dundun contributes a repertoire-speci c timeline and hook pattern. The
ensemble is led by Jeli Madi Kuyate, a senior percussionist retired from the state-sponsored Ballet National
du Mali. Internationally acclaimed drummer Drisa Kone plays the second jembe, and their long-time
ensemble co-performer Madu Jakite plays the dundun. These three professionals were apprenticed to the
same master, and have performed together since the 1970s. The video features the opening 30 seconds of a
recording of a piece called ‘Wolosodòn’ or ‘Jòndòn’, which means ‘serfs’ dance’ or ‘captives’ dance’,
relating to the subaltern social strata of unfree folk in the feudal precolonial past. ‘Wolosodòn’ is part of the
core repertoire of standard pieces for jembe music in Bamako, the capital of Mali. In what follows, I provide
a brief analysis that includes information on subdivision timing ratios.

The lead drummer, Kuyate, opens the piece by establishing the second jembe’s accompaniment pattern, a C14.P9
long–short ostinato with a timbral melody spanning two beats. Open tones and bass tones in alternation
mark the onbeat positions, while a slap on each upbeat adds timbral accent to this metric position and thus
gives anacrustic drive to the rhythm. On average, the long and the short element in each beat relate by a
ratio of 57:43 (percentage of the beat duration) or 4:3. There is only minimal variation to this average, as the
individual ratios fall within a very narrow range (56:44–59:41). It is a straightforward and conventional way
for a lead drummer to call up a piece by taking on the accompanist’s pattern. The second jembe player, Drisa
Kone, therefore almost immediately (near the end of Cycle 1) takes over ‘his’ pattern from the lead
drummer. He adopts the timbre sequence and rhythmic gure as well as Kuyate’s timing pro le with
remarkable precision: Kone’s mean subdivision ratio on the second jembe is also 57:43, and the variability
in his timing throughout the excerpt is as small as Kuyate’s; the standard deviation from the mean value
amounts to no more than about 1%.

Immediately after the dundun’s entry in Cycle 2, the ensemble synchronization appears very tight. In Cycle C14.P10
3, lead drummer Kuyate starts soloing by unfolding a line of slaps, still in a lightly swung two-element
rhythm very close to the accompaniment. This choice of a still super-simple rhythm speaks of an
unpretentious attitude, of containment and control; many other players would have chosen some more
complex rhythm here. Yet this does not mean that Kuyate refrains from sophistication. He plays all slaps in
that series with his stronger, right hand and uses the left one to gently mu e the membrane in between the
note onsets, thus softly dampening the harmonics and muting the decay of each slap. The e ect is
somewhat comparable to an electric guitarist’s pumping the wah-wah pedal in time with an eighth note
beat. It is a soft e ect, yet of considerable appeal among jembe players; in the specialists’ discourse and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
performance practice, it is iconic of old-school playing and associated with the liveliness of a ‘breathing’ or
p. 259 ‘speaking’ jembe sound. Kuyate adds to the expressivity of the phrase by attening the swing ratio to
54:46, which he will later repeat with pinpoint exactitude when he comes back to the same motive in a
slightly longer phrase, in Cycles 8 and 9. The timing thus emerges to represent a systematic treatment of
the respective phrase. The local attening of the swing ratio lets the upbeat strokes deviate only minimally
from the prototype of 57:43, amounting to di erences in the concerned durations of around (a hardly
perceptible) 15 ms. As minuscule as this contrast appears, it adds some nuance to the anacrustic momentum
of the rhythm.

In the rst half of Cycle 4, lead drummer Kuyate lets up on his drum for a second and speaks out in honour C14.P11
of his late jembe master (Yamadu Dunbia, 1917–2002), who was of Woloso descent himself and had played
‘Wolosodòn’ as his signature piece; the other players nod their heads in consent. In the second halves of
Cycles 4 and 5, respectively, Kuyate puts out a motif which for the rst time in this performance manifestly
realizes a ‘triplet feel’ or, ternary subdivision. His ternary rhythm still sounds fully consonant with the
second jembe’s ostinato and the dundun’s hook, and this is because his triplet subdivisions also show a
periodic pattern of non-isochronous subdivision, namely short–medium–long, with a ratio of about
22:35:43 on average. The lead drummer’s third triplet stroke and the accompanist’s shu e second stroke
receive the same percentage of the beat-span (43%), and the strokes occur with a high degree of
simultaneity. In Cycles 6 and 7, Kuyate presents two turns of a phrase which features triplets similar to
those of Cycles 4 and 5, which he combines with an extra stroke towards the end of each cycle suggesting a
roll, i.e. a rapid ornamental guration beyond metric perception. In Cycles 8 and 9, the lead-drummer
comes back to a two-element long–short rhythm almost identical to the one played earlier in Cycle 3.

Listening Options Based on Categorical Isochrony C14.S3

The timing of the beat subdivisions in Video Sample 1 poses problems to conventional theoretical C14.P12
perspectives. When listened to in isolation, a two-element rhythm with a ratio of 57:43 can perhaps be
heard as an expressive variation of a duplet rhythm based on an isochronous binary subdivision. However,
this listening would clash with the frequent insertion of triplets; in particular, it could not account for the
smoothness with which triplets appear to be embedded in the polyrhythmic surface of ‘Wolosodòn’.

Alternatively, the accompaniment ostinato may be heard as a shu e rhythm, in which the two sounds are C14.P13
perceived as mapped onto the rst and third elements of an underlying isochronous ternary subdivision of
the metric beat (see Figure 14.2). While this accounts for the insertion of triplets, the 57:43 performance
timing of the second jembe’s ostinato would still represent a massive deviation from this perspective’s 2:1
(67:33) expectation. Moreover, according to categorical rhythm perception theory, the 57:43 timing, which
falls right midway between 1:1 (50:50) and 2:1 (67:33) and thus can be assumed to lie near the boundary
between the two categories, should induce a strong degree of perceptual ambiguity or con ict in the
listener. Studies of African American styles of popular music have suggested exactly this when they
p. 260 described comparable subdivision timings as ‘playing between the cracks’ (Doleac 2013) or ‘nervous’ and
‘hybrid’ (Stewart 2000). In fact, I felt such ambiguity myself when listening to ‘Wolosodòn’ and comparable
rhythms in Mali during my initial years of engaging with this music. Today, I feel no ambiguity or tension
when listening to ‘Wolosodòn’—it sounds perfectly consonant and organic to my ears. Seemingly, at some
stage of my career I learned to perceive the piece di erently according to how my own playing was evaluated
by my masters in Mali and my students in Germany. It appears that much of this enculturation occurred
during my one-year apprenticeship in Bamako in 1997–8, when I performed the role of accompanist on the
second jembe several times per week at weddings, spirit possession ceremonies, and other social and ritual
occasions, playing simple ostinatos, always loud and often fast, for hours.

Figure 14.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Shu le rhythm in Western sta notation (top tier) mapped to a metric cycle (bottom row) of four beats (second row from bottom) C14.F2
with ternary subdivision (third row from bottom).

The subjective listening experience set out above is consistent with more objective long-term ethnographic C14.P14
observations I have made in Mali. For instance, I have frequently observed that non-isochronous
subdivision timing patterns were performed not only by skilled instrumentalists but also by the many young
children who frequently sit close to dance celebrations and imitate the specialists by playing on toy drums.
Importantly, I have never seen any performance of a piece based on non-isochronous subdivisions fall back
to an isochronous version, as one might expect in situations that demanded reduced complexity. For
example, there are situations where an ensemble of novice apprentices shows di culties in ensemble
synchronization or a teacher struggles to convey a speci c subdivision timing pattern to European students.
I witnessed plenty of such cases, yet actually never heard any Malian drummer perform ‘Wolosodòn’ based
on anything close to a plain shu e and evenly spaced triplets, which would be the simplest and safest, most
regular thing to do from a Western isochronous metre perspective (though not necessarily the most
exciting!). (Audio samples 1-A/B provided at the accompanying website demonstrate how a typical phrase
for ‘Wolosodòn’ would actually sound when based on isochronous versus non-isochronous subdivision
grids.)

To conclude, it seems implausible to conceive of a performance timing pattern such as the 57:43 (4:3) ratio C14.P15
in the accompaniment for ‘Wolosodòn’ in Video sample 1 as a deviation from an isochronous reference
structure, because the hypothetical isochronous prototype in this case would represent an extreme outlier
to the actual performance timings. While it is not necessary for the prototype of a category to lie exactly
p. 261 midway in the respective segment of the perceptual space, it can be expected to lie somewhere in its more
intensely used interior, but not at its unpopulated fringes.
An Alternative Model Allowing for Categorical Non-isochrony C14.S4

In light of these problems with listenings based on categorical isochrony, I propose the alternative C14.P16
hypothesis: that a complex ratio such as 57:43 (4:3), falling between the categories of 2:1 and 1:1, can
constitute a distinct rhythmic category of its own. I further hypothesize that non-isochronous beat
subdivisions related by such complex ratios, if consistently organized in recurrent patterns, can serve as
metric temporal reference structures. Figure 14.3 shows how this hypothetical separate category would fall

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
in a range of ratios where Western participants display a signi cant gap indicative of perceptual ambiguity
at the border between the 1:1 and 2:1 prototypes. The ratios in this range have been conceived of as irregular
and thus di cult and problematic, if not explicitly unmusical, in European and Euro-American theories of
2
music since antiquity.

Figure 14.3

Upper tier: Schematic representation of the perceptual categorization of the space of two-element rhythms according to the C14.F3
current state of research (repeated from Figure 14.1). The lower tier indicates the alternative hypothesis of an additional
category with a prototype at about 4:3.

Adapted from Polak et al. (2018: fig. 1).

A metric type based on a rhythmic prototype such as 4:3 at the subdivision level would challenge some C14.P17
fundamental assumptions widely held in metric theory. Figure 14.4 illustrates this novelty by comparing
hierarchic tree representations of four di erent metric types. Figures 14.4a and 14.4b represent two
common types of isochronous metre which are both based on a four-beat cycle. The type in 14.4a is quali ed
by two layers of binary beat subdivision (4/4 in Western sta notation), whereas the type in 14.4b has one
layer of ternary subdivision (12/8). A listening of ‘Wolosodòn’ (Video Sample 1) based on categorical
isochrony, as sceptically discussed above, would suggest the latter metric type (12/8). In these two metric
types, pulse-trains at both the beat and the subdivision levels consist of isochronous elements throughout.
Consequently, the pattern of bifurcation in the hierarchical structure is strictly periodic; each pulsation at
p. 262 the same
p. 263

metric level is subdivided by the same number of pulsations at the faster level. Standard accounts of metric
theory suggest that these qualities represent essential preconditions for metric pickup and stable
coordination.
Figure 14.4

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023

Schematic illustrations of hierarchic trees of selected metric types: (a) Isochronous 4-beat metric cycle with two layers of C14.F4
isochronous subdivision, binary and quaternary, (b) isochronous 4-beat cycle with one level of isochronous ternary subdivision,
(c) non-isochronous 3-beat cycle (short–short–long) with one level of isochronous 7-element subdivision, (d) isochronous 4-beat
cycle with two levels of non-isochronous binary (long–short) and non-isochronous ternary subdivision (short–medium–long).

By contrast, the metric types in Figure 14.4c and 14.4d include non-isochronous pulse-streams. Figure 4c C14.P18
outlines a non-isochronous three-beat cycle with a seven-element subdivision, typically notated as 7/8.
This metric type is popular in music from the Balkans among other places, and was used as a prime example
in the aforementioned studies of non-isochronous beat cycles (Goldberg 2017; Hannon et al. 2012). Finally,
Figure 14.4d illustrates my hypothesis of non-isochronous metre in music from Mali in the example of a
metric cycle of four isochronous beats with two levels of non-isochronous subdivision, one binary (long–
short), and the other ternary (short–medium–long). This is how I hear ‘Wolosodòn’ (Video Sample 1), for
instance.

In the two non-isochronous metric types (C and D), the structure of hierarchic coordination of pulse-trains C14.P19
varies in the course of the cycle. In type C, the rst and second beats are bifurcated into two subdivisions,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
whereas the third one hosts three; in type D, the rst pulse at the rst subdivision level is bifurcated into
two subdivisions at the second level, whereas the second pulse at the rst level remains un-bifurcated.
While these patterns appear simple, they are complex when compared to the strict periodicity of
hierarchical coordination in isochronous metres. This complexity notwithstanding, the layers in the non-
isochronous metric hierarchies (C and D) are as fully nested into each other as they are in the isochronous
types (A and B).

Empirical Evidence C14.S5

Arguably, the hypothesis of non-isochronous subdivision based metre illustrated in Figure 14.4d could C14.P20
elegantly explain the performance practice exempli ed in Video Sample 1. However, the hypothesis
contradicts the assumptions of strict isochrony and periodic bifurcation as fundamental preconditions of
metre perception. I have therefore engaged in a series of ve empirical research projects to provide evidence
that would enable an evaluation of the hypothesis. This section o ers concise summaries of the key ndings
of these studies. The order is chronological, yet at the same time displays a developing trajectory of
methodologies, from performance-timing analysis to controlled laboratory experiments, from qualitatively
grounded and exploratory to quantitative and hypothesis-testing approaches, and from individual to
collaborative research work ows.

Study 1: Performance Timings of Jembe Drumming from Mali C14.S6

Study 1 comprised an empirical performance-timing analysis based on a corpus of multi-track live C14.P21
p. 264 ensemble recordings, an approach applied for the very rst time to music from Mali (or indeed anywhere
else in Africa). Over the course of eldwork conducted in 2006 and 2007, I hired ve di erent ensembles of
experienced jembe and dundun players for open-air studio sessions both in Bamako and in two villages in
the rural area to the south of the capital. Using a portable digital multi-track machine, we recorded about
180 takes of over 20 pieces of repertoire. A clip-on microphone attached to each drum created a separate
audio track of each ensemble member’s playing, which involved only a little acoustical crosstalk from the
neighbouring instruments and thus allowed precise identi cation of the onset timings of the drum-strokes
constituting the composite ensemble polyrhythms.

In all analysed performances, the beat-span turned out to be isochronous. Three cardinalities of beat C14.P22
subdivision were found in the repertoire (binary, ternary, and quaternary), and each of these subdivision
types occurred with stable non-isochronous timing patterns in most of the pieces. The relation between
subdivision timing pattern and repertoire emerged very clearly: for most pieces, the pattern of subdivision
timing proved invariant, as if it were ‘written’ into the composition. The in-depth analysis of several
recordings of a piece called ‘Manjanin’, which features a pattern of non-isochronous ternary beat
subdivision (short–medium–long), demonstrated that such timing patterns can occur with an amazing
degree of consistency. It proved stable across ve parameters, namely:

1. di erent recordings by di erent rural and urban ensembles; C14.P23


2. di erent recordings of the same ensembles; C14.P24

3. di erent instruments (jembe versus dundun) involving di erent playing techniques and motor skills; C14.P25

4. di erent musical roles (lead, timeline and accompaniment) involving di erent phrase types and C14.P26
rhythmic structures;

5. a wide range of tempos. (Polak 2010) C14.P27

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Study 2: Performance Timings of Other Styles of Drumming from Mali C14.S7

The second project took up the methodology developed in Study 1 and expanded the scope from jembe music C14.P28
to two other traditions of ensemble drumming—the distinct drum ensembles, repertoires, and styles of two
ethnolinguistic groups from central and western Mali (Bamana and Khasonka, respectively). In the two
respective corpuses of recordings made in 2012, about half the pieces consistently showed non-isochronous
subdivision timings, and the patterns and the degree of consistency with which they were performed turned
out to be very similar to those found in jembe music. Moreover, detailed analyses of two pieces suggested
that non-isochronous timing patterns can occur simultaneously in two nested layers of beat subdivision.
One of those two pieces nests a binary long–short pattern into a ternary short–medium–long subdivision,
p. 265 very much like the jembe piece ‘Wolosodòn’ exempli ed in Video Sample 1 and modelled in Figure 14.2.
The nesting of both levels of subdivision turned out to be highly systematic and almost perfectly accurate,
meaning that the second binary and the third ternary subdivision onsets consistently occur nearly
simultaneously. The other piece analysed in detail entails a rst layer of ternary subdivision timed
according to a long–short–short pattern and a second, quaternary layer patterned short–short–long–long.
Here, too, the second level of subdivision is constructed through the further bifurcation of only one element
at the rst level—the long one (Polak and London 2014). In other words, the quality of non-isochrony in
some pieces of Malian drum ensemble music is recursive across metric levels.

Study 3: Timing Variability and Ensemble Synchronization in Jembe Drumming C14.S8

The third project was a data-intensive, computational repeat study of 15 recordings selected from the C14.P29
corpus made in the context of Study 1. In contrast to Studies 1 and 2, which relied on analysing the onset
timings of a limited number of typical phrases, in Study 3 we extracted the onset timing for each drum
stroke of all players in each of the 15 recordings in the corpus, amounting to about 40,000 time-points, each
representing the onset of an individual drum stroke. This allowed us to con rm on statistical grounds the
key ndings of Studies 1 and 2, namely the near-perfect isochrony of the beat-span in all pieces and the
high degree of consistency and stability in the subdivision timing patterns for each individual piece. The
variability of the timings in the two non-isochronous pieces in the corpus (‘Wolosodòn’ and ‘Manjanin’)
proved not to be signi cantly higher than in the third, isochronous piece that we had included for
comparison. Moreover, the analysis of microrhythmic asynchronies amongst the onsets of di erent
ensemble members in the same metric positions indicated that the quality of ensemble synchronization was
extremely tight in all three analysed pieces. Importantly, we did not nd a signi cant di erence between
the isochronous and non-isochronous pieces in this respect (Polak et al. 2016).

Study 4: A Listener Test of Subdivision Timing Patterns in Jembe Music C14.S9

The fourth project involves experimental listener tests. We selected two typical phrases from recordings of C14.P30
the jembe piece ‘Manjanin’ (cf. Studies 1 and 3), and created realistically sounding versions of these phrases
based on manipulated subdivision timing patterns, including the original one (short–medium–long) as
well as isochronous and inverted ones (long–medium–short).
p. 266 We presented hundreds of pairings of two (identical or di erent) manipulations, and asked drummers and C14.P31
dancers in Bamako, who had much experience of the piece, (1) whether the two versions in a pair were, rst,
the same or di erent, and (2) which version was the better representation of the piece. There were no
di erences between the drummer and the dancer groups, and both responded identically. Participants
reliably recognized almost all manipulations as same or di erent, but performed quite poorly at
discrimination between versions based on the patterns short–medium–long and short–long–long. In the
preference task, the version based on the original subdivision timing pattern (short–medium–long)

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
together with the perceptually identical one (short–long–long) gained the highest score, whereas other
manipulations were ranked increasingly lower as their di erence from the original became more
pronounced. For example, the version based on an isochronous subdivision earned considerably lower
ratings than the original, and the version based on a reversed sequence of non-isochronous elements
(long–medium–short) scored substantially less. In summary, our groups of enculturated participants
displayed an aesthetic ideal of preference for a subdivision pattern modelled after the style-speci c
performance practice (Neuho et al. 2017).

Study 5: A cross-cultural listener test of context-free simple rhythms C14.S10

Study 5 used the experimental paradigm of nger-tapping, presenting simple rhythms without melodic or C14.P32
other context to participants whose task was to tap along to the given rhythms on a little percussion
instrument in best possible synchrony. We tested jembe drummers in Mali, musicians (mostly
percussionists specializing in Western art music) in Germany, and folk musicians as well as musicians
trained in Western art music in Bulgaria. At a slow tempo, where the pacing rhythms could be expected to be
perceived as representing the metric beat level, we replicated key ndings of previous research (see Figure
14.1). All participant groups reliably reproduced the 1:1 ratio but, in the case of uneven rhythms, showed a
strong bias toward the 2:1 ratio; they softened a 3:1 pacing stimulus and sharpened a 3:2 stimulus, involving
a distortion in the direction of 2:1 in both cases. However, the Bulgarian folk musicians responded to the 3:2
stimulus with a slightly smaller degree of distortion than the other groups, which correlates with the fact
that 3:2 rhythms feature prominently in some genres of Bulgarian folk-dance music at tempos such as the
one used in the stimuli.

At a fast tempo, which a ords perception of the pacing rhythms as subdivisions of an isochronous beat, we C14.P33
found that all tested groups exactly reproduced a 1:1 stimulus and performed quite well on a 2:1 rhythm, but
strongly diverged in their response to a 58:42 (approximately 4:3) stimulus. The German and both Bulgarian
groups massively distorted this complex ratio towards 2:1, whereas the Malians reproduced it with near-
p. 267 perfect delity. This suggests that the Malian musicians—but not the Bulgarian and German musicians
—have a perceptual rhythmic category whose prototype lies in the vicinity of 4:3 (Polak et al. 2018).
Summary and discussion C14.S11

This chapter has evaluated the hypothesis that in the context of drum ensemble music from Mali, non- C14.P34
isochronous beat subdivision patterns constitute a structural aspect of rhythm and metre rather than
representing an expressive timing deviation from these temporal reference structures, as established
theories of rhythm and metre in music suggest. I have presented various kinds of empirical evidence. First,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
performance timing analyses of various styles of dance drumming from Mali suggested that the
performance of isochronous and non-isochronous subdivision timing patterns alike is characterized by
qualities of consistency, stability and intersubjective coherence (Studies 1–3). This is inconsistent with the
assumption that 1:1 and 2:1 represent the only available perceptual prototypes. The latter assumption would
predict that the non-isochronous patterns in question, which fall about midway 1:1 and 2:1, will induce
perceptual ambiguity, structural instability, and increased degrees of variability, exibility, and
contingency in the performance timings. Our performance timing analyses systematically belie this
prediction. Secondly, we found that patterns of non-isochronous subdivision can occur recursively across
two smoothly nested layers of subdivision, where the second layer is constructed from the bifurcation of
exactly one element in the rst one—the longest one (Study 2). This suggests that the pattern of uneven
subdivision timing in one layer is used as a perceptual reference framework for the timing of the other. A
third type of evidence came from two psychological experiments that tested how musicians in Mali perceive
non-isochronous beat subdivision patterns. Study 4 suggested that the same rhythmic gures based on
di erent subdivision timing patterns involve structurally di erent percepts, and found a clear aesthetic
preference for the original patterns that are used in the speci c repertoire’s performance practice. However,
while the original performance timing patterns consist of three di erent durational elements (e.g. short–
medium–long), the experiment suggested that the medium category is hardly perceptible, and the
subdivision structure thus can be constructed from only two durational classes in the listeners’ perception
(short–long–long). In other words, whereas subdivision timing patterns in general are perceptually and
aesthetically salient, not all aspects of the timing patterns that can be measured in performances contribute
to this relevance; for instance, the medium category of the short–medium–long pattern that can be found
in many performances does not. Study 5 demonstrated that Malian musicians have a separate perceptual
prototype for categorical rhythm perception at approximately 4:3, i.e. exactly midway between 1:1 and 2:1.
The ndings of Studies 4 and 5 directly contradict important counter-arguments against the hypothesis,
namely, that subdivision timing patterns based on complex ratios are either perceptually irrelevant or
p. 268 substantially detrimental to the perception of rhythmic and metric structure due to universal cognitive
constraints.

To summarize, there is diverse and robust evidence in support of the hypothesis that uneven subdivision C14.P35
patterns can constitute rhythmic and metric reference structures. Demonstrably, these patterns do not
appear irregular in the perception and evaluation of enculturated listeners. Let me end the chapter by
considering some of the implications of this nding for our broader understanding of how our perceptual
capacities are shaped by the cultural environment in which we grow up and dwell. Study 5 (Polak et al. 2018)
revealed two important quali cations regarding the nding that a somewhat complex ratio such as 4:3 can
constitute a perceptual prototype for categorical rhythm perception. First, only musicians in Mali, who
prominently use ostinato rhythms based on the same ratio in their style- and repertoire-speci c
performance practice, appear to have this category; other musicians, who have been socialized in musical
cultures that do not prominently use the respective ratio, do not show the category, regardless of the
musicians’ expertise or the sophistication of their musical culture. Secondly, the Malian group displayed the
category only in response to stimuli that were presented in the fast tempo typical of the beat subdivisions
that feature the 4:3 ratio in their performance practice, and not in response to stimuli presented at a slower
tempo that is typical of beat rates rather than of subdivision rates. Taken together, those two correlations
between style-speci c performance practice and perceptual prototypes suggest a strong role of the music-
cultural environment in framing links between performance and perception prototypes. Evidently, the
Malian musicians’ perceptual capacity for non-isochronous subdivision-based metre is not a cognitive
universal; our research suggests that instead they have developed this speci c capacity in their music-
cultural environment.

This nding aligns with recent propositions that the temporal reference framework within which rhythmic C14.P36
structure is constructed, and from which expressive deviations are perceived, is constituted not primarily by

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
symbolic systems (e.g. a score) nor by universal cognitive constraints, but rather by statistical inference
from actual performances of repertoires and styles frequently practised and perceived in one’s music-
cultural environment (Honing 2013; Leech-Wilkinson 2009; Pearce 2018; van der Weij 2020; van der Weij et
al. 2017). More generally, this nding is also consistent with views from the ecological theory of perception
on perceptual learning (Gibson 1963; 1969; Gibson and Gibson 1955; Goldstone 1998; Goldstone and
Hendrickson 2010; Goldstone et al. 2010; Goldstone et al. 2015), as well as with the concept of ‘perceptual
narrowing’, which describes processes of environment-dependent unlearning of generic perceptual
capacities.

An exemplary case is speech perception. While very young infants can distinguish speech sounds which are C14.P37
not necessarily used in their mother tongue, the capacity for categorical phoneme perception by older
infants, children, and adults increasingly focuses on those phonemes which are used in the mother tongue,
to the cognitive detriment of other phonemes (Maurer and Werker 2014; Werker 1995). Erin Hannon and
Sandra Trehub have found equally uent perception of isochronous and non-isochronous beat cycles by
p. 269 very young Western infants, yet a perceptual privileging of isochronous beats is observable from the age
of one year and tends to increasingly stabilize with age (Hannon and Trehub 2005a; 2005b). Adults’
perceptual capacity to uently process non-isochronous metre seems to depend on its being used in musical
performance and listening practices (Hannon et al. 2012), and it becomes impoverished in the context of
musical cultures which do not frequently use non-isochronous pulse-streams. Importantly, this
demonstrates that the perceptual bias towards isochrony found in Western listener groups is due to their
speci c enculturation rather than a universal cognitive predisposition (Hannon 2010), as much as the Malian
participants’ special perceptual capacity for non-isochronous subdivision is the result of an increasing
ability to perceptually di erentiate the diverse stimuli that their music-cultural environment has to o er.
This is inconsistent with the nativist assumption that a bias towards isochrony is culturally universal
because of a biological constraint on human cognition. The wide spread of this assumption may instead
indicate a Western cultural bias in scholarly and scienti c accounts of rhythm and metre.

Acknowledgements C14.S12

I would never have been able to develop the argument as summarized in this chapter without the amazing C14.P38
amount of support I am happy to acknowledge here. The research, including ve eldwork stays in Mali,
was generously funded by the German Research Council (Deutsche Forschungsgemeinschaft, DFG) in three
phases summing up to a total of more than six years; another year was nanced by the University for Music
and Dance Cologne. I have pro ted tremendously from discussions with collaborators and colleagues,
including Andreas Meyer, Hans Neuho , Timo Fischinger, Anne Danielsen, Tellef Kvifte, Mats Johansson,
Dan Goldberg, Luis Jure, Martín Rocamora, Andre Holzapfel, David Locke, and Rich Jankowsky. In
particular, collaborating with Justin London and Nori Jacoby has pushed my research beyond what I could
conceive possible. I am also extremely grateful to the anonymous reviewers who invested their time and
energy into improving my research proposals and research papers, and to Mark Do man, who o ered
invaluable editorial suggestions on how to improve the quality of the present chapter. Finally, since 1991,
my hosts in Mali, including drummer Madu Jakite and linguists Mohammed Diallo and Salabary Doumbia
and their families, proved more than indispensable for my eldwork in this country. Thank you all.
Accompanying media files C14.S13

Video sample 1.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
0:00

[Video/Audio] available online

C14.P39
(The video is accessible on the handbookʼs accompanying website Trio performance of Woloso. Jeli Madi Kuyate: first
p. 270 jembe (stereo position: centre); Madu Jakite: dundun (le ; out of sight); Drisa Kone: second jembe (right). Recorded in
Bamako, January 2006, by the author. The recording is primed by two artificial clicks giving the beat tempo before the first
jembe starts on Beat 3 of Cycle 1.

Audio sample 1-A and 1-B.

[Video/Audio] available online

Synthesized renditions of a typical phrase for ʻWolosodònʼ, modeled a er the triplety phrase played in Cycles 4 and 5 of Video C14.P40
sample 1. The two versions were constructed by Nori Jacoby from single-note instrument sounds played and recorded by the
author; both are quantized to a metric grid, which entails that they do not contain timing variations within the ensemble parts
nor asynchronies amongst ensemble parts. They thus allow for a fair comparison of the subdivision timings, which are non-
isochronous (23:35:42) in 1-A and isochronous (33% per element) in 1-B.

[Video/Audio] available online

Notes

1. The concepts of ʻbeatʼ and ʻpulseʼ largely overlap. I term ʻpulseʼ any stream of felt metric pulsations, independently of C14.N1
which layer in the metric hierarchy this pulse would represent. By contrast, I reserve the concept of beat for metric pulse-
streams that constitute the middle level in a hierarchy of three metric layers (cycle, beat, subdivision), whereas the beat
subdivision represents the lowest/fastest layer in that hierarchy.

2. I am grateful to David Cohen (pers. comm., 18 Sept. 2018) for providing me with information on Aristoxenus of Tarentum C14.N2
addressing the ratios falling between 2:1 and 1:1 as ʻirrationalʼ or ʻarrhythmicʼ in his fragment Elementa rhythmica (4th
century BC ).
References C14.S14

Agawu, K. (2006). Structural analysis or cultural analysis? Competing perspectives on the ʻstandard patternʼ of West African C14.P41
rhythm. Journal of the American Musicological Society 59(1): 1–46. http://www.jstor.org/stable/10.1525/jams.2006.59.1.1
Google Scholar WorldCat

Arom, S. (1984). Structuration du temps dans les musiques dʼAfrique centrale. Revue de Musicologie 70(1): 5–36. C14.P42

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Benadon, F. (2006). Slicing the beat: Jazz eight-notes as expressive micro-timing. Ethnomusicology 50(1): 73–98. C14.P43
Google Scholar WorldCat

Burns, J. M. (2010). Rhythmic archetypes in instrumental music from Africa and the diaspora. Music Theory Online 16(4). C14.P44
http://www.mtosmt.org/issues/mto.10.16.4/mto.10.16.4.burns.html
Google Scholar WorldCat

Butterfield, M. W. (2011). Why do jazz musicians swing their eighth notes? Music Theory Spectrum 33(1): 3–26. C14.P45
http://www.jstor.org/stable/10.1525/mts.2011.33.1.3
Google Scholar WorldCat

Clarke, E. F. (1985). Structure and expression in rhythmic performance. In P. Howell, R. West, and I. Cross (eds), Musical structure C14.P46
and cognition, 209–236. Academic Press.
Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (1987). Categorical rhythm perception. An ecological perspective. In A. Gabrielsson (ed.), Action and perception in C14.P47
rhythm and music, 19–33. Royal Swedish Academy of Music.
Google Scholar Google Preview WorldCat COPAC

Clarke, E. F. (1999). Rhythm and timing in music. In D. Deutsch (ed.), The psychology of music (2nd edn), 473–500. Academic C14.P48
Press.
Google Scholar Google Preview WorldCat COPAC

p. 271 Clarke, E. F. (2000). Categorical rhythm perception and event perception. Proceedings of the International Music Perception and C14.P49
Cognition Conference, Keele University. https://www.escom.org/proceedings/ICMPC2000/Tue/Clarke.htm
WorldCat

Clarke, E. F., and Do man, M. (2014). Expressive performance in contemporary concert music. In D. Fabian, R. Timmers, and C14.P50
E. Schubert (eds), Expressiveness in music performance: Empirical approaches across styles and cultures, 98–114. Oxford
University Press.
Google Scholar Google Preview WorldCat COPAC

Doleac, B. (2013). Strictly second line: Funk, jazz, and the New Orleans beat. Ethnomusicology Review 18, 1–16.
Google Scholar WorldCat

Drake, C., and El Heni, J. B. (2003). Synchronizing with music: Intercultural di erences. Annals of the New York Academy of C14.P52
Sciences 999(1): 429–437.
Google Scholar WorldCat

Essens, P., and Povel, D.-J. (1985). Metrical and non-metrical representations of temporal patterns. Perception and Psychophysics C14.P53
37: 1–7.
Google Scholar WorldCat

Feldman, N. H., Gri iths, T. L., and Morgan, J. L. (2009). The influence of categories on perception: explaining the perceptual C14.P54
magnet e ect as optimal statistical inference. Psychological Review 116(4): 752–782. https://doi.org/10.1037/a0017196
Google Scholar WorldCat

Fitch, W. T. (2012). The biology and evolution of rhythm: unravelling a paradox. In P. Rebuschat, M. Rohmeier, J.A. Hawkins, and C14.P55
I. Cross (eds), Language and music as cognitive systems, 73–95. Oxford University Press.
https://doi.org/10.1093/acprof:oso/9780199553426.003.0009
Google Scholar Google Preview WorldCat COPAC

Fitch, W. T. (2013). Rhythmic cognition in humans and animals: Distinguishing metre and pulse perception. Frontiers in Systems C14.P56

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Neuroscience 7. https://doi.org/10.3389/fnsys.2013.00068
Google Scholar WorldCat

Fraisse, P. (1956). Les structures rhythmiques. Publications universitaires de Louvain. C14.P57


Google Scholar Google Preview WorldCat COPAC

Fraisse, P. (1982). Rhythm and tempo. In D. Deutsch (ed.), The psychology of music, 149–180. Academic Press. C14.P58
Google Scholar Google Preview WorldCat COPAC

Gibson, E. J. (1963). Perceptual learning. Annual Review of Psychology 14: 29–56. https://doi.org/ C14.P59
Google Scholar WorldCat

Gibson, E. J. (1969). Principles of perceptual learning and development. Appleton-Century-Cro s. C14.P60


Google Scholar Google Preview WorldCat COPAC

Gibson, J. J., and Gibson, E. J. (1955). Perceptual learning: Di erentiation or enrichment? Psychological Review 62(1): 32–41. C14.P61
https://doi.org/10.1037/h0048826
Google Scholar WorldCat

Goldberg, D. (2017). Bulgarian metre in performance. Doctoral dissertation, Yale University. C14.P62
Google Scholar Google Preview WorldCat COPAC

Goldstone, R. L. (1998). Perceptual learning. Annual Review of Psychology 49: 585–612. C14.P63
https://doi.org/10.1146/annurev.psych.49.1.585
Google Scholar WorldCat

Goldstone, R. L., and Hendrickson, A. T. (2010). Categorical perception. Wiley Interdisciplinary Reviews: Cognitive Science 1(1): 69– C14.P64
78. https://doi.org/10.1002/wcs.26
Google Scholar WorldCat

Goldstone, R. L., Landy, D. H., and Son, J. Y. (2010). The education of perception. Topics in Cognitive Science 2(2): 265–284. C14.P65
https://doi.org/10.1111/j.1756-8765.2009.01055.x
Google Scholar WorldCat

Goldstone, R. L., de Leeuw, J. R., and Landy, D. H. (2015). Fitting perception in and to cognition. Cognition 135: 24–29. C14.P66
https://doi.org/10.1016/j.cognition.2014.11.027
Google Scholar WorldCat

Hannon, E. E. (2010). Musical enculturation: How young listeners construct musical knowledge through perceptual experience. In C14.P67
S. P. Johnson (ed.), Neoconstructivism: The New Science of Cognitive Development, 132–156. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Hannon, E. E., and Trehub, S. E. (2005a). Metrical categories in infancy and adulthood. Psychological Science 16(1): 48–55. C14.P68
https://doi.org/10.1111/j.0956-7976.2005.00779.x
Google Scholar WorldCat

Hannon, E. E., and Trehub, S. E. (2005b). Tuning in to musical rhythms: Infants learn more readily than adults. Proceedings of the C14.P69
National Academy of Sciences of the United States of America 102(35): 12639–12643. https://doi.org/10.1073/pnas.0504254102
Google Scholar WorldCat

Hannon, E. E., Soley, G., and Ullal-Gupta, S. (2012). Familiarity overrides complexity in rhythm perception: A cross-cultural C14.P70
p. 272 comparison of American and Turkish listeners. Journal of Experimental Psychology: Human Perception and Performance
38(3): 543–548. https://doi.org/10.1037/a0027225
WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Harnad, S. R. (1990). Categorical perception: The groundwork of cognition. Cambridge University Press. C14.P71
Google Scholar Google Preview WorldCat COPAC

Honing, H. (2013). Structure and interpretation of rhythm in music. In D. Deutsch (ed.), The psychology of music (3rd edn), 369– C14.P72
404. Academic.
Google Scholar Google Preview WorldCat COPAC

Honing, H., and de Haas, W. B. (2008). Swing once more: Relating timing and tempo in expert jazz drumming. Music Perception C14.P73
25(5): 471–476.
Google Scholar WorldCat

Hood, M. (1971). The Ethnomusicologist. McGraw Hill. C14.P74


Google Scholar Google Preview WorldCat COPAC

Jacoby, N., and McDermott, J. H. (2017). Integer ratio priors on musical rhythm revealed cross-culturally by iterated C14.P75
reproduction. Current Biology 27(3): 359–370. https://doi.org/10.1016/j.cub.2016.12.031
Google Scholar WorldCat

Johansson, M. (2009). Rhythm into style: Studying asymmetrical grooves in Norwegian folk music. Doctoral dissertation, University C14.P76
of Oslo.

Johansson, M. (2017). Non-isochronous musical metres: Towards a multidimensional model. Ethnomusicology 61(1): 31–51. C14.P77
https://doi.org/10.5406/ethnomusicology.61.1.0031
Google Scholar WorldCat

Jones, M. R., and Boltz, M. (1989). Dynamic attending and responses to time. Psychological Review 96(3): 459–491. C14.P78
Google Scholar WorldCat

Kalender, B., Trehub, S. E., and Schellenberg, E. G. (2013). Cross-cultural di erences in metre perception. Psychological Research C14.P79
77(2): 196–203.
Google Scholar WorldCat

Keil, C. (1987). Participatory discrepancies and the power of music. Cultural Anthropology 2(3): 275–283. C14.P80
https://doi.org/10.1525/can.1987.2.3.02a00010
Google Scholar WorldCat

Koetting, J. (1970). Analysis and notation of West African drum ensemble music. Selected Reports in Ethnomusicology 1(3): 116– C14.P81
146.
Google Scholar WorldCat

Kubik, G. (Ed.). (1988). Zum Verstehen afrikanischer Musik: Ausgewählte. Reclam. C14.P82
Google Scholar Google Preview WorldCat COPAC

Kvi e, T. (2007). Categories and timing: On the perception of metre. Ethnomusicology 51(1): 64–84. C14.P83
Google Scholar WorldCat

Large, E. W. (2008). Resonating to musical rhythm: theory and experiment. In S. Grondin (ed.), The psychology of time, 189–232. C14.P84
Emerald.
Google Scholar Google Preview WorldCat COPAC

Large, E. W., and Jones, M. R. (1999). The dynamics of attending: How people track time-varying events. Psychological Review C14.P85
106(1): 119–159.
Google Scholar WorldCat

Large, E. W., and Kolen, J. F. (1994). Resonance and the perception of musical metre. Connection Science 6: 177–208. C14.P86

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Large, E. W., and Snyder, J. S. (2009). Pulse and metre as neural resonance. Annals of the New York Academy of Sciences 1169: 46– C14.P87
57. https://doi.org/10.1111/j.1749-6632.2009.04550.x
Google Scholar WorldCat

Leech-Wilkinson, D. (2009). The changing sound of music: Approaches to studying recorded musical performances. Centre for the C14.P88
History and Analysis of Recorded Music.
Google Scholar Google Preview WorldCat COPAC

Lerdahl, F., and Jackendo , R. (1983). A generative theory of tonal music. MIT Press. C14.P89
Google Scholar Google Preview WorldCat COPAC

Locke, D. (1982). Principles of o beat timing and cross-rhythm in southern Ewe dance drumming. Ethnomusicology 26(2): 217– C14.P90
246.
Google Scholar WorldCat

Locke, D. (2010). Yewevu in the metric matrix. Music Theory Online 16(4). C14.P91
https://www.mtosmt.org/issues/mto.10.16.4/mto.10.16.4.locke.html
Google Scholar WorldCat

London, J. (2012). Hearing in time: Psychological aspects of musical metre (2nd edn). Oxford University Press. C14.P92
Google Scholar Google Preview WorldCat COPAC

Maurer, D., and Werker, J. F. (2014). Perceptual narrowing during infancy: A comparison of language and faces. Developmental C14.P93
Psychobiology 56(2): 154–178.
Google Scholar WorldCat

p. 273 Merker, B., Madison, G., and Eckerdal, P. (2009). On the role and origin of isochrony in human rhythmic entrainment. Cortex C14.P94
45(1): 4–17. https://doi.org/10.1016/j.cortex.2008.06.011
Google Scholar WorldCat

Neuho , H., Polak, R., and Fischinger, T. (2017). Perception and evaluation of timing patterns in drum ensemble music from Mali. C14.P95
Music Perception 34(4): 438–451. https://doi.org/10.1525/mp.2017.34.4.438
Google Scholar WorldCat

Nketia, J. H. K. (1974). The music of Africa. Norton. C14.P96


Google Scholar Google Preview WorldCat COPAC

Pearce, M. T. (2018). Statistical learning and probabilistic prediction in music cognition: Mechanisms of stylistic enculturation. C14.P97
Annals of the New York Academy of Sciences 1423(1): 378–395. https://doi.org/10.1111/nyas.13654
WorldCat

Polak, R. (2010). Rhythmic feel as metre: Non-isochronous beat subdivision in jembe music from Mali. Music Theory Online 16(4). C14.P98
https://www.mtosmt.org/issues/mto.10.16.4/mto.10.16.4.polak.html
WorldCat
Polak, R., Jacoby, N., Fischinger, T., Goldberg, D., Holzapfel, A., and London, J. (2018). Rhythmic prototypes across cultures: A C14.P99
comparative study of tapping synchronization. Music Perception 36(1): 1–23.
Google Scholar WorldCat

Polak, R., Jacoby, N., and London, J. (2016). Both isochronous and non-isochronous metrical subdivision a ord precise and C14.P100
stable ensemble entrainment: A corpus study of Malian jembe drumming. Frontiers in Neuroscience 10: 285.
https;//doi.org/10.3389/fnins.2016.00285
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Polak, R., and London, J. (2014). Timing and metre in Mande drumming from Mali. Music Theory Online 20(1). C14.P101
https://www.mtosmt.org/issues/mto.14.20.1/toc.20.1.html
WorldCat

Povel, D.-J. (1981). Internal representation of simple temporal patterns. Journal of Experimental Psychology: Human Perception C14.P102
and Performance 7: 3–18.
Google Scholar WorldCat

Prögler, J. A. (1995). Searching for swing: Participatory discrepancies in the jazz rhythm section. Ethnomusicology 39(1): 21–54. C14.P103
Google Scholar WorldCat

Ravignani, A., Delgado, T., and Kirby, S. (2016). Musical evolution in the lab exhibits rhythmic universals. Nature Human C14.P104
Behaviour 1(1): 7. https://doi.org/10.1038/s41562-016-0007
Google Scholar WorldCat

Ravignani, A., and Madison, G. (2017). The paradox of isochrony in the evolution of human rhythm. Frontiers in Psychology 8: C14.P105
1820. https://doi.org/10.3389/fpsyg.2017.01820
Google Scholar WorldCat

Repp, B. H. (1984). Categorical perception: Issues, methods, findings. Speech and Language: Advances in Basic Research and C14.P106
Practice 10: 243–335.
Google Scholar WorldCat

Repp, B. H., London, J., and Keller, P. E. (2005). Production and synchronization of uneven rhythms at fast tempi. Music C14.P107
Perception 23(1): 61–78. https://doi.org/10.1525/mp.2005.23.1.61
Google Scholar WorldCat

Repp, B. H., London, J., and Keller, P. E. (2011). Perception–production relationships and phase correction in synchronization C14.P108
with two-interval rhythms. Psychological Research 75(3): 227–242. https://doi.org/10.1007/s00426-010-0301-8
Google Scholar WorldCat

Repp, B. H., London, J., and Keller, P. E. (2012). Distortions in reproduction of two-interval rhythms: When the ʻattractor ratioʼ is C14.P109
not exactly 1:2. Music Perception 30(2): 205–223. https://doi.org/10.1525/mp.2012.30.2.205
Google Scholar WorldCat

Savage, P. E., Brown, S., Sakai, E., and Currie, T. E. (2015). Statistical universals reveal the structures and functions of human C14.P110
music. Proceedings of the National Academy of Sciences 112(29): 8987–8992. https://doi.org/10.1073/pnas.1414495112
Google Scholar WorldCat

Stewart, A. (2000). New Orleans, James Brown, and the rhythmic transformation of American popular music. Popular Music, 19
(3), 293–318.
Google Scholar WorldCat

Tal, I., Large, E. W., Rabinovitch, E., Wei, Y., Schroeder, C. E., Poeppel, D., and Zion Golumbic, E. (2017). Neural entrainment to the C14.P111
beat: The ʻmissing-pulseʼ phenomenon. Journal of Neuroscience 37(26): 6331–6341. https://doi.org/10.1523/JNEUROSCI.2500-
16.2017
Google Scholar WorldCat

p. 274 van der Weij, B. (2020). Experienced listeners: Modeling the influence of long-term musical exposure on rhythm perception. Institute C14.P112
for Logic, Language and Computation.
Google Scholar Google Preview WorldCat COPAC

van der Weij, B., Pearce, M. T., and Honing, H. (2017). A probabilistic model of metre perception: simulating enculturation. C14.P113
Frontiers in Psychology 8: 824. https://doi.org/10.3389/fpsyg.2017.00824

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353469971 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Waterman, R. (1952). African influence on the music of the Americas. In S. Tax (ed.), Acculturation in the Americas, 207–218. C14.P114
University of Chicago Press.
Google Scholar Google Preview WorldCat COPAC

Werker, J. F. (1995). Exploring developmental changes in cross-language speech perception. In L. R. Gleitman and M. Liberman C14.P115
(eds), An invitation to cognitive science, vol. 1: Language (2nd edn), 87–106. MIT Press.
Google Scholar Google Preview WorldCat COPAC

Wesolowski, B. C. (2015). Timing deviations in jazz performance: The relationships of selected musical variables on horizontal C14.P116
and vertical timing relations. A case study. Psychology of Music 44(1): 75–94. https://doi.org/10.1177/0305735614555790
Google Scholar WorldCat

Yates, C. M., Justus, T., Atalay, N. B., Mert, N., and Trehub, S. E. (2017). E ects of musical training and culture on metre perception. C14.P117
Psychology of Music 45(2): 231–245. https://doi.org/10.1177/0305735616657407
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
CHAPTER

15 Towards a Cognitively
C15 Based Quanti cation of Metrical
Dissonance 
Mark Gotham

https://doi.org/10.1093/oxfordhb/9780190947279.013.12 Pages 277–C15.P143


Published: 08 December 2021

Abstract
Metrical dissonance is a powerful tool for creating and manipulating musical tension. The relative
extent of tension can be more or less acute depending (in part) on the type of dissonance used and
moving among those dissonance types can contribute to the shape of a musical work. This chapter sets
out a model for quantifying relative dissonance that incorporates experimentally substantiated

principles of cognitive science. Online supplements for this chapter at www.oup.com/us/ohtm


provide mathematical formalisations of the principles discussed (Supplement 1) and an interactive
guide for testing out these ideas (Supplement 2). We begin with a basic model of metre where a
metrical position’s weight is given simply by the number of pulse levels coinciding there. This alone
enables a telling categorization of displacement dissonances for simple metres and a rst sense of the
relative di erences between them. These arbitrary weighting ‘values’ are then re ned on the basis of
tempo and pulse salience. This provides a more subtle set of gradations that re ect the cognitive
experience of metre somewhat better while still retaining a clear sense of the simple principles that
govern relative dissonance. Additionally, this chapter sees the model applied in a brief, illustrative
analysis and in a preliminary extension to ‘mixed’ metres (5s, 7s,…). This sheds light on known
problems such as the relative stability of mixed metres in di erent rotations, and suggests a new way
of thinking about mixed metres’ relative susceptibility to metrical dissonance.

Keywords: metre, rhythm, metrical dissonance, pulse salience, mixed metre, theory, analysis
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
An Introduction to Metre and Dissonance C15.S1

‘DISSONANCE ’ is one of those multi-purpose terms with a range of uses both within and beyond musical C15.P1
analysis. Outside of music, terms such as ‘cultural’ or ‘cognitive dissonance’ usually have negative
connotations, but music has a way of making a virtue of the ‘wrong’ or ‘ill- tting’. Musical dissonance is
not ‘bad’ in the same sense as some of these other contexts, though canons of musical ‘rules’ often require

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
‘proper’ handling and resolution, at least in the case of tonal counterpoint.

Within music, the separation of ‘consonances’ (octaves, fths…) from ‘dissonances’ (semi-tones, C15.P2
tritones…) in the realm of pitch is well established. The broad structure of this consonant/dissonant divide
(or continuum) is robust, though theorists have long argued over the details, such as: the need for further
divisions (e.g. separating ‘perfect’ from ‘imperfect’ consonances); the membership of those categories
(particularly in relation to the perfect fourth); and how this system came to be (naturally or arti cially).

The overtone series is frequently invoked as a ‘natural’ basis for the origin-of-consonance story. C15.P3
Notwithstanding essential di erences between pitch and frequency (with the former as a mental construct
based on—but not equivalent to—the latter), frequency ratios which reduce to simple fractions (and thus
appear towards the start of the overtone series) are usually asserted to be the most consonant: the octave
1
(2:1), followed by the fth (2:3) and so on. By contrast, the combination of frequencies with closer absolute
p. 278 frequency values (separated by a semi-tone, for instance) yields a harmonic interval further up the
overtone series which is more dissonant than the combination of pitches whose frequencies, though less
2
absolutely similar, are in a simpler proportion.

This suggests that we need to distinguish between two kinds of similarity: the ‘proportional’ similarity of C15.P4
simple fractions and consonant ratios, versus the ‘proximal’ similarity of close absolute values and
dissonant intervals. Apart from musical contexts, ‘clashing’ colours provide a useful analogy for the latter
correspondence of proximity with dissonance. Among the criteria that go into combining colours in
aesthetically pleasing ways is the avoidance of ‘too-similar’ combinations such as red with pink.

Metrical Structure and Metrical Dissonance C15.S2

More recently (in the history of music theory, that is), these same notions of consonance and dissonance C15.P5
3
have been applied to metre. The modern literature on ‘metrical dissonance’ was galvanised by Krebs
(1999), who in turn cites a Berlioz article of 1837 as the earliest explicit comparison using these terms (pp.
4
13, 16). Notwithstanding notable di erences, there are clear grounds for this analogy between harmonic
and metrical dissonance and su ciently strong commonalities to set them out in comparable ways, once
again based on the simplicity of proportional relationships.

Figure 15.1 provides a schematic representation of musical metre in terms of layered pulse levels that is C15.P6
5
similar to many such diagrams in the literature theorizing Western classical metre. There are pulse streams
for: quavers (at the bottom), crotchets (immediately above, coinciding with every other quaver), minims
(every other crotchet), and semi-breves (every other minim). This also sets out the idea of metrical
hierarchy as a many-levelled system of strong–weak patterns: the more pulse levels coinciding at a point in
the metrical cycle, the more strongly weighted it is in the hierarchy. In Figure 15.1, each pair of neighbouring
levels is in a 2:1 relationship. Let’s call this structure a ‘binary’ metre.
Figure 15.1

A schematic diagram for metrical structure in terms of pulse layers. C15.F1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
In any such binary metre, the most strongly weighted position is at the start of the whole span (the bar or C15.P7
higher-level unit), and this is followed by a position at the bisection of that span (which is as far as possible
from the strongest position), and so on. Time signatures like 4/4, 2/4, and 2/2 all relate to this type of
structure. By including one or more level of 3-grouping, we can make the alternative structures
corresponding to 6/8, 3/4, and indeed all of the metres of the common practice period. For any metre in this
p. 279 system, the positions immediately beside those of the greatest metrical weight are particularly weak. In
short, the distinction between ‘proportional’ and ‘proximal’ similarity is relevant here too.

In straightforward cases, musical and notational cues align unambiguously and we can be con dent of ‘the’ C15.P8
metre. Metrical dissonance can arise where those implications diverge. As London has it: ‘Metric
dissonances occur when secondary accents […] undermine the established metre to the point that a
secondary metric framework may emerge’ (London n.d.). If there’s a part of the music that you can readily
hear in terms of a metre other than the prevailing one, you have metrical dissonance. This divergence may
be established by a part, perhaps just one musical parameter (like chord changes), by performance cues, and
6
even by the notation alone.

Metrical dissonance has recently become a focal topic for much musical analysis of common practice, tonal C15.P9
7 8
music, and popular–vernacular styles. As such, there is a clear disciplinary motivation to make the
theoretical apparatus as robust as possible. That theoretical toolkit includes several ‘metrical spaces’ (or
Zeitnetze) analogous to the well-known ‘tonal spaces’ (Tonnetze) which help to map out the structural
9
options; this sometimes extends to mapping out relations between metrical dissonance types but almost
10
never to ranking the relative strength of those dissonances.

While Krebs is right to note that ‘it is by no means obvious precisely how the intensity of dissonances should C15.P10
be ranked’ (1999: 57) and even that ‘of more musical signi cance than inherent intensity is contextual
intensity’ (p. 58), there is much to be gained by formalizing intuitive principles to see what factors emerge,
and how pervasively. This chapter seeks to provide the groundwork for such a venture, combining cognitive
and theoretic approaches in the service of a kind of musical understanding relevant to practitioners and
analysts alike. The more deeply we understand how metrical dissonance works (both in theory and in
practice), the richer our appreciation will be, and the more informed and versatile we can be in discussing,
performing, and creating music featuring this device.

Types of Metrical Dissonance C15.S3

Figure 15.2

Simple prototypes for displacement and grouping dissonance. C15.F2


Figure 15.2 sets out the two main types of metrical dissonance: ‘displacement’ and ‘grouping’. Figure 15.2a C15.P11
sees two simple rhythms in a ‘displacement’ dissonance. All else being equal, hearing those parts separately
would likely lead to di erent assumptions about ‘the’ metre. The lowest part aligns with the notated metre.
The upper part can be understood in terms of a version of the ‘same’ metre, but one which has been
p. 280 ‘displaced’ by a crotchet, potentially inverting assumptions about which beats are strong and weak. We’ll
focus primarily on displacement dissonance in this chapter, because it is the simpler of the two for the
present purposes.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Figure 15.3 provides an example from Bill Withers’ iconic ‘Lean on Me’ (1972). This contrasting section (at C15.P12
1’45” and 2’36”, to the lyrics ‘you just call…’) introduces a kind of rhythmic ingenuity that can be helpful to
consider in terms of displacement dissonance. While the ‘main’ metre and grouping is clear from the
foregoing context and the vocal part, other parts of the texture shift to consistently displaced positions.
Starting with the bass guitar, while the anacrustic pattern in the low register stays with the main metre, the
bass has a prominent note in a higher register on beat 3 that asserts itself as a di erent voice with an
alternative take on the metrical phase, potentially displacing the metre by a half-cycle to beat 3. The drum
kit, in turn, emphasizes beats 2 and 4 with its back-beat pattern. This can be heard in terms of displacement
by a quarter-cycle, though it is important to note that this is a particularly listener-dependent case: many
listeners will nd this pattern so familiar that the very presence of that displaced emphasis serves instead to
indicate where the strong beats ‘really’ are. Finally, and most strikingly, there are hand claps (prominent in
the mix) on the 2nd and 6th quavers of the bar, suggesting an eighth-cycle displacement. The ‘real’ phase
of the metre may never be in doubt, but the rhythmic-metrical richness of this passage can be helpfully
understood in terms of the systematic syncopations that potentially point to alternative displacements of
the metre.

Figure 15.3

Displacement dissonance in Bill Withersʼ ʻLean on Meʼ (authorʼs transcription). Brackets indicate the displacements discussed in C15.F3
the text.

‘Grouping’ dissonance sees the combination of two metres with di erent internal organization. Figures C15.P13
15.2b and 15.2c set out a ‘grouping dissonance’: the upper part implies an internally consistent grouping in
3s (perhaps indicating a 6/8 time signature), which contradicts the implications of the lower part’s
grouping in 4s (2/4). In Figure 15.2b, it is the lower, ‘2/4’ part that aligns with the notated metre; in Figure
15.2c, the ‘same’ dissonant pairing is notated to t the upper part’s ‘6/8’.

Figure 15.2’s 3-against-4 rhythm is probably the most ubiquitous form of rhythmic-metrical patterns that C15.P14
11
p. 281 can be readily heard in terms of ‘grouping dissonance’. Figure 15.2 sets out complete cycles of this
grouping dissonance, but partial cycles are more common in musical practice. Versions of this pattern are
common in popular musics from classic jazz (such as the melodic rhythm of Gershwin’s ‘I Got Rhythm’) to
electronic dance music (EDM), and owes its origins to the Afro-Cuban tresillo rhythm which can be viewed
as a partial cycle of the 3-against-4 dissonance broadly equivalent to the rst two bars in Figure 15.2b.
Figure 15.4 sets out an example of the partial cycle in EDM from Calvin Harris’ ‘I’m Not Alone’ (2009). Once
again, the ‘real’ metre is clear from the context, and this ri (along with much of the rhythmic content of
this example) sets out a con icting pattern based on grouping in 3s. The upper part of Figure 15.4 provides
the ri while the lower part makes explicit the continuing 3-grouping, the possible reading of this rhythm
in terms of 12/16 (with brackets), and the point at which this con ict is re-set (‘!’) which leads to a re-set at
the start of the next note and bar.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Figure 15.4

Grouping dissonance in Calvin Harrisʼ ʻIʼm Not Aloneʼ (2009) as an example of a popular 3 against 4 pattern (transcribed by the C15.F4
author).

Scope and Psychology C15.S4

Before we embark on working through a model for quantifying these dissonances, a few words on the scope C15.P15
of this chapter and the psychology of metre are in order. The quantitative model advanced here simply seeks
to address musical contexts which simultaneously use patterns that imply di erent metrical structures. The
model seeks to integrate some apparently robust results emerging from recent work in musical cognition,
but it certainly does not purport to solve all of the many, important cognitive questions at play.

For instance, the notion of ‘combining’ metres is far from straightforward. If a metre is a kind of grid C15.P16
against which we parse rhythmic events, what does it mean to have two at the same time? Don’t you have to
choose between them, as in the well-known rabbit/duck image of gestalt psychology? The psychological
literature does indeed suggest that ‘there is no such thing as [a perceptual mechanism for] polymeter’
12
(London 2012: 50), but that instead listeners tend to hear one viable metre in the context of another. We
may switch between options, but not hold both simultaneously. We distinguish between a ‘ gure’ and a
‘ground’, and speak of a ‘ gure–ground reversal’ for swapping between those options, though that is all
consistent with the two metrical options being viable metres. In any case, this model is neutral on that
question: it quanti es the approximate ‘amount’ of dissonance, without speci cally committing to the
‘true’ metre or any more detailed perceptual rami cations.

p. 282 Further, this theoretical continuum from consonant to dissonant is certain to be more complex in C15.P17
perceptual practice. For instance, there is likely to be a point at which dissonance becomes so acute it can no
longer be registered in quite the same way. This would be consistent with (and related to) the ‘trill
threshold’: a speed beyond which alternation between di erent entities becomes too fast to register those
13
entities separately, and so they fuse. Again, this model does not resolve such questions or set a benchmark
for a ‘metrical dissonance threshold’, though the model’s integration of pulse salience does go some way
towards that goal and may provide a useful framework for developing and testing ideas of that kind.

Quantifying Displacement Dissonances Discretely C15.S5

To quantify dissonance, we rst need a quanti cation of metrical weighting, converting ‘strong, weak’ into C15.P18
numerical values like ‘2, 1’. The section on ‘Tempo-sensitive, Pulse-salience Weightings’ sets out a
relatively new, cognitively based method for doing this; but we will start with the simple discrete values
more common in music theory. What these accounts share is a notion of ‘levels’ in metrical structure and
the attribution of a value per level.
Proportional and Proximal Similarity Revisited C15.S6

Theories of metre linked to representations like Figure 15.1 tend to see metrical weighting in terms of the C15.P19
number of pulses levels coinciding at each position in the cycle. On that basis, the ‘strong, weak, medium,
weak’ of a simple 4/4 becomes 3,1,2,1 and adding an extra pulse level expands this to 4,1,2,1,3,1,2,1. In 4/4,
this later case could equate to one-bar cycle, with the quaver level weighted at 1, the crotchet level at 2
(crotchets and quavers coinciding), the minim level 3, and the semi-breve 4.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Setting this metre against itself in displacement by a half-cycle (so with the two metres in anti-phase) C15.P20
yields the situation shown in (a):

(a) C15.P21

Original 4 1 2 1 3 1 2 1 Sum Mean

Half-cycle displacement 3 1 2 1 4 1 2 1

Resulting di erence 1 0 0 0 1 0 0 0 2 0.25

The representation in (a) summarizes the location of the di erences, and can also be made to give a broad C15.P22
sense of the relative strength of dissonance, for instance with the sum and/or average values provided on
p. 283 the right. For instance, by almost any mathematical reckoning, this anti-phase arrangement yields
‘little’ dissonance. Displacement by a quarter-cycle increases the overall dissonance further (b).

(b) C15.P23

Original 4 1 2 1 3 1 2 1 Sum Mean

Quarter-cycle displacement 2 1 4 1 2 1 3 1

Resulting di erence 2 0 2 0 1 0 1 0 6 0.75

Finally, displacement by a single unit (eighth-cycle) yields the strongest dissonance of all (c). C15.P24

(c) C15.P25

Original 4 1 2 1 3 1 2 1 Sum Mean

Unit displacement 1 4 1 2 1 3 1 2

Resulting di erence 3 3 1 1 2 2 1 1 14 1.75

This gives us a mathematical way of expressing the di erence between proportional and proximal similarity C15.P26
already discussed. Displacement by simpler proportions such as a half-cycle create little dissonance;
displacement by a lesser distance (and less simple proportion, such as an eighth-cycle) yield stronger
dissonance. It all comes down to the di erences between the metrical positions set in con ict.
On Cycle Length and the Arithmetic Mean C15.S7

The summative counts of di erences between metres may work as a heuristic for simple cases, but we are C15.P27
interested here in comparing types of metrical dissonance, including those of di erent cycle length. For
instance, adding a metrical level to our example metre (with either a semi-quaver or whole note pulse)
would double the cycle length and slightly more than double the summative dissonance values. That doesn’t
accord with musical experience. It is primarily the displacement which causes dissonance, not the presence

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
or absence of an additional pulse level. Taking an average allows us to compare like with like—in this case,
adding only a small amount to the dissonance values in each case.

Here we use the simplest option for that averaging: the arithmetic mean. This gives us a single value that C15.P28
represents the ‘average di erence per position’ in a complete cycle. We continue to use arbitrary units; the
relevant comparison is between comparable, relative values (types of dissonance).

p. 284 In taking a single ‘average’ value to stand for the metrical dissonance as a whole, we are greatly reducing C15.P29
the overall information. There will certainly be times we wish to study the ‘raw’ data of the full dissonance
pro le (including positive and negative values), but we need some reduction of this kind if we are to rank
relative dissonance, and the di erences between them. A single number is the simplest choice, and the
14
arithmetic mean is the only standard distance metric suitable to this data.

Categories of Displacement Dissonances Based on Metrical Levels C15.S8

We can already establish some formalisations and generalisations to get a sense of how these displacement C15.P30
dissonances compare. Let L be the length of the metre measured by the total number of units at the lowest
level under consideration and Dl be the metrical dissonance value associated with displacement by l units. In
the examples above, L = 8 and the displacement types are D4, D2, and D1.

First, note that we are not (at least not yet) considering one metre in terms of the other, but simply C15.P31
comparing the two on equal terms. As such, D depends only on the distance between the two metres (not
which came rst), and so we have an equivalence between positive and negative displacements (D+ l = D-l).

Secondly, it is not usually meaningful to talk of displacement by more than a full cycle, because there is C15.P32
always an equivalent metrical comparison within the metre. Therefore, we invoke modular addition to
establish that DL + l = D+ l. This is somewhat analogous to removing octaves from the question of intervals: in
some contexts, we need to speak speci cally of a ‘major 10th’, but often the generic ‘major 3rd’ is more
appropriate.

Our third equivalence is more musically interesting. Apart from the fact that D+ 1 = D-1, it is also true of the C15.P33
present model that D1 = D3 for displacement of binary metres. This is because the dissonances involve the
same set of direct comparisons between representatives of the metrical levels involved. The table in (d) sets
out D3, showing equivalent comparisons and values to D1 (shown above).

(d) C15.P34

Original 4 1 2 1 3 1 2 1 Sum Mean

D+3 1 2 1 4 1 2 1 3

Resulting di erence 3 1 1 3 2 1 1 2 14 1.75


This illustrates the fact that di erent displacement distances can give rise to the same dissonance value. On C15.P35
this basis, we can de ne categories of displacements which have the same dissonance value. These are
p. 285 linked to the metrical levels involved: each displacement belonging to a single category involves the same
set of comparisons between representatives of the metrical levels (as shown for D1 = D3).

In the case of our 8-unit binary metre, there are four distinct metrical levels, and so three nontrivial classes C15.P36
of displacement dissonance for this metre (ways of comparing those distinct levels). The membership of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
these categories is as follows:

1. D+∕−1 and D+∕−3 (as discussed), as well as D+∕−5 (= D−∕+3), and D+∕−7 (= D−∕+1); C15.P37

2. D+∕−2 and D+∕−6; C15.P38

3. D+∕−4 remains unique: there is only one half-cycle. C15.P39

Clearly this categorical equivalence between D1, D3, D5, and D7 means that the system does not (yet) account C15.P40
for the di erentiation of metrical positions. This is addressed in the section titled ‘Position- and context-
speci city’, along with a discussion of musical context. Nevertheless, the categorization that emerges from
the model in this state is a useful initial overview which is in line with commonly used representations of
metre and which provides a formalization of idea of that displacements which relocate strong metrical
positions (high levels) to weak ones (low levels) yield strong dissonance.

Equivalent categories of this kind emerge for any metre based on duple or triple grouping between each C15.P41
metrical level. That accounts for all ‘common practice’ metres (3/4, 6/8,…metres like 5/4 and 7/8 are
discussed below in ‘Mixed Metres: Metrical Weighting’). The same level-based equivalences hold, but not
for the same values. For instance, in a 6/8 metre with three levels in the pulse ratio 1:3:6, (cycle length = 6),
D1≠ D3 any more, but rather D1 = D2 = D4 = D5. Table 15.1 shows this correspondence. As for our binary metre,
the location of the di erences changes, but the overall dissonance value is equivalent for all members of the
category, and is greater than ‘simpler’ displacements involving only the higher levels: here again,
displacement by a half cycle returns a much lower dissonance value (this time of 1/6).

Summary of the Discrete-Weighting Model C15.S9

To summarize the model advanced so far: assuming widely adopted music-theoretic de nitions of Western C15.P42
classical metre, the following signi cant generalizations can be made for displacement dissonance among
all common practice metres (4/4, 2/4, 3/4, 6/8…):

• The number of distinct, non-trivial categories of equivalent metrical dissonance is one fewer than the C15.P43
number of metrical levels.

• The least dissonant categories involve changes among only the highest levels; the most dissonant C15.P44
categories involve changes at all levels (usually displacement by one unit, D1).

p. 286 • Not only do the dissonance values increase as one proceeds from the highest to the lowest levels, C15.P45
but the di erences between those successive dissonance values also seem to increase. This property
holds for almost all plausible values, though it does not quite generalize across the full tempo range
when we include the tempo-weighted salience mode below (see Online Supplement 1
www.oup.com/us/ohtm for a full explanation).

• Adding a new, highest metrical level (e.g. making ‘2/4’ into ‘2/2’) adds to the dissonance values for C15.P46
each category, and does so equally by the same value to each dissonance category, thus preserving
these hierarchies.
Table 15.1 The four equivalent displacements of a 6/8 metre in the ratio 1:3:6 C15.T1

D1

Original 3 1 1 2 1 1

Displacement 1 3 1 1 2 1 Sum Mean

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Di erence 2 2 0 1 1 0 6 1

D2

Original 3 1 1 2 1 1

Displacement 1 1 3 1 1 2 Sum Mean

Di erence 2 0 2 1 0 1 6 1

D4

Original 3 1 1 2 1 1

Displacement 1 2 1 1 3 1 Sum Mean

Di erence 2 1 0 1 2 0 6 1

D5

Original 3 1 1 2 1 1

Displacement 1 1 2 1 1 3 Sum Mean

Di erence 2 0 1 1 0 2 6 1

Online Supplement 1 www.oup.com/us/ohtm explains these claims further and provides mathematical C15.P47
formalisations of the membership of the categories and the di erences between levels. The rest of this
section pauses the theorizing to explore these ideas in some historical and musical contexts. The following
section, ‘Improvements to the model’, then advances the metrical dissonance model, beginning by
incorporating the promised tempo-sensitive weighting of metrical salience. This introduces meaningfully
weighted values for the rst time and thus incorporates some signi cant aspects of musical perception;
p. 287 remarkably, it does so without great changes to the outcomes of the discrete heuristic developed so far—
most of the results continue to hold. Subsequent discussion concerns the e ect of individuating metrical
positions and the relative status of mixed metres (5s, 7s—see ‘Mixed metres: metrical weighting’).

This basic form of the model connects with an idea which might be summarized by the maxim: ‘The lower C15.P48
the level, the greater the change.’ As I have discussed elsewhere (Gotham 2015a; 2015b; 2017), irregularity
appears to exist more readily at higher metrical levels than at the most salient, tactus levels. All else being
equal, it is much more disruptive to insert an additional sub-tactus pulse unit (turning 2/4 into 5/8, for
instance) than to insert an extra bar in the phrase. The di erence may be visualized in terms of a marching
soldier having to either (1) negotiate unequal beat patterns (perform a kind of ‘Aksak’ dance), or (2) simply
perform one extra left–right cycle.

That being the case, a slightly playful (and more than slightly reductive) grand narrative presents itself. In C15.P49
common practice tonal music, metre tends to x the notated levels (of pulse, tactus, and bar) in simple,
15
regular pulse streams, and to enjoy more exible, irregular patterns at higher, hypermetrical levels. Then,
among the major developments of the extended common practice (and around the same time as
Schoenberg’s (1951) ‘emancipation of the dissonance’), we start to see those irregular patterns migrate to
the tactus level. Perhaps then we should speak not only of an ‘emancipation of the dissonance’, but also of
its metrical parallel: the ‘emancipation of the tactus levels’.

As always with such reductive accounts founded on analogies, there are many problems. First, it seems to C15.P50
imply that mixed metres were ‘invented’ in the extended common practice, whereas ‘discovered’ would be a

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
more apt term: for instance, Bartók was a major protagonist of their introduction to Western art music, yet
this was a result of his exposure to their long-standing and highly sophisticated use in the folk musics of
many Balkan regions. Second, ‘early’ or ‘pre-common practice era’ music undermines this trajectory by
virtue of the dizzying array of rhythmic-metrical complexities in isorhythmic motets and the like. Finally,
even the idea that metrical dissonance within common practice music followed an historical trajectory of
usage from high to low metrical levels has many more exceptions than needed to ‘prove the rule’. The rst
movement of Bach’s Sixth Brandenburg Concerto provides a ready example.

Analytical Interlude: Bachʼs Brandenburg Concerto No. 6, Movement 1 C15.S10

This section o ers a short analysis which simultaneously provides an illustration of how the model as C15.P51
outlined so far may be musically useful, while also motivating the need to nesse the model further. The
rst movement of Bach’s Sixth Brandenburg Concerto is so full of metrical dissonance (and harmonic
16
p. 288 stasis) that it is tempting to describe it even in proto-minimalist terms. There is canonic/imitative
writing in almost every bar (excepting only rare moments with one solo melodic part playing) and thus,
insofar as we expect the same melodic material to recur with the same metrical identity, then we also have
metrical dissonance throughout this movement.

Five types of melodic material are treated canonically in this movement: C15.P52

• Ritornello idea; imitative at the distance of a quaver in the viola ‘da braccio’ parts. C15.P53

• Episode idea 1; at the minim; Violoncello and all violas. C15.P54

• Episode idea 2; at the minim; in the violas da braccio. C15.P55

• Episode accompanimental idea; at the crotchet; in the violas da gamba. C15.P56

• Episode idea 3; at the minim; between the Violoncello and the violas da gamba. C15.P57

These are shown in musical notation in Figure 15.5, and their usage falls neatly into categories by section as C15.P58
shown in Table 15.2.

Naturally for a ritornello form movement, the main distinction is between ritornellos and episodes. C15.P59
However, in that normative practice, we might expect those ritornellos to stand as the relatively stable
sections in comparison with the more developmental episodes. In fact, it is in the ritornellos that we nd the
most erce metrical dissonance, with imitation at the distance of a quaver (D2). In the episode sections, by
contrast, the canonic imitations are at the much less striking intervals of the crotchet (D4) and minim (D8).

The minim dissonance is particularly weak, so much so that there is a strong case for considering ‘metrical C15.P60
17
dissonance’ meaningless at the minim in this movement. The time signature is alla breve, but signi cant
events occur about as frequently at the half-bar as they do on the ‘downbeat’. These include not just the
chord changes (which are rare enough), but also ritornellos. Even the nal (‘recapitulatory’) ritornello in
the tonic key commences at the mid-point of bar 114, leading the nal cadence to also fall mid-bar.
Arguably, then, the movement may be better considered without metrical levels above the minim-pulse, in
terms of a kind of ‘2/4’. In any case, as long as we include or exclude that level consistently, it does not
a ect the internal comparisons. This was one of the key points included in ‘Summary of the discrete-
weighting model; and reinforces the need for averaging.

What may be more analytically signi cant is the relative di erence between metrical dissonance types. C15.P61
Recalling that the di erence between types increases at the faster unit levels, we may consider making a
case for further di erentiation of the ritornellos (D2 only) from the episode sections (D4 and D8) on the basis
of the increased gap between dissonance types (D2 − D4> D4 − D8). Before making any claim so speci cally

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
quantitative, we would need a much more robustly numerical system than that o ered by model in its
p. 289 current form. The following section goes some way towards providing that more usable quanti cation.

Figure 15.5

Musical incipits for the canonic imitations in the first movement of Bachʼs Brandenburg Concerto No. 6 (the first appearances of C15.F5
each).
p. 290 Table 15.2 Displacement dissonance in the first movement of Bachʼs Sixth Brandenburg Concerto, section by section. Dx C15.T2
indicates the distance between successive entries in semi-quaver units

Bar Material Distance Dissonance

1 Ritornello Quaver D2

17 Episode idea 1 Minim D8

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
21 Episode idea 2 Minim D8

Episode accompanimental idea Crotchet D4

25 Ritornello Quaver D2

28 Episode idea 1 Minim D8

31 Episode idea 3 Minim D8

32 Episode idea 1 Minim D8

36 Episode idea 2 Minim D8

Episode accompanimental idea Crotchet D4

40 (Solo)

46 Ritornello Quaver D2

53 (Solo)

56 Episode idea 1 Minim D8

62 Episode idea 2 Minim D8

Episode accompanimental idea Crotchet D4

65 Episode idea 1 Minim D8

68 Episode idea 3 Minim D8

69 Episode idea 2 Minim D8

73 Ritornello Quaver D2

80 Episode idea 3 Minim D8

84 (Solo)

86 Ritornello Quaver D2

92 Episode idea 1 Minim D8

96 Episode idea 2 Minim D8

Episode accompanimental idea Crotchet D4

101-2 (Ritornello) Quaver D2


102 Episode idea 1 Minim D8

107 Episode idea 2 Minim D8

Episode accompanimental idea Crotchet D4

110 Episode idea 2 (no acc.) Minim D8

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
114 Ritornello Quaver D2

p. 291
Improvements to the Model C15.S11

Tempo-Sensitive, Pulse-Salience weightings C15.S12

The model developed so far helps make sense of metrical hierarchies and the logical consequences that C15.P62
emerge from them for the relative ‘strength’ of di erent metrical dissonance types. The main limitation to
this as a quantitative theory is the equal weighting of each metrical level.

The literature suggests a way to improve upon this situation is by quantifying the respective contribution of C15.P63
each metrical level. This would endow this model with more meaningful values and also account somewhat
better for the psychology of metrical listening. The broad consensus of studies like Parncutt (1994) and van
Noorden and Moelants (1999) suggests that pulses may be more or less easy to parse metrically: that there
is a central, ‘easiest’ range around 100bpm and this tapers o as pulses become simply too fast/slow. For
instance, we seem to nd it much easier to maintain accurately and independently a pulse rate of 1 second
than of 10 seconds. The term ‘salience’ broadly captures this notion of relative ease in mentally processing
pulses at di erent rates.

In Gotham (2015a), I developed this relatively well-established principle into a more speculative means of C15.P64
using those individual pulse salience weightings to account for salience at the level of metres (which
comprise many pulse levels). To illustrate this, let’s reprise from that paper the example of a 6-level binary
metre with crotchet MM = 100bpm. Table 15.3 gives the salience values for each level individually and for the
cumulative weightings that accrue as they are added together. The cycle of metrical weights for each
successive position that emerges from this approach is shown in Figure 15.6.

Table 15.3 Salience values for the example binary metre C15.T3

Level a b c d e f

Pulse stream (IOI in seconds) 0.15 0.3 0.6 1.2 2.4 4.8

Individual pulse salience 0.133 0.604 1 0.604 0.133 0.011

Cumulative (metrical) salience 0.133 0.738 1.738 2.342 2.476 2.487

This addition of metrical levels works in exactly the same way for the weighted version as in the discrete C15.P65
p. 292 case—the system still simply adds a value for each metrical level present. The systems di er in how
much they add per level. The discrete system weighted every level’s contribution equally (at a value of 1) and
returned overall metrical weighting values of 1, 2, 3…; in the salience-weighted case, we weight each level
with a salience function. So if x is the length of the unit pulse in seconds, and the salience value for that level
is S(x), then the lowest metrical weighting is given by S(x) alone, and further levels add the corresponding
salience values for each additional level. For the binary metre example in Table 15.3, the salience values for
each level (a, b, c, d, e, f) are given by a = S(0.15), b = a + S(0.3), c = b + S(0.6), d = c + S(1.2), e = d + S(2.4), f = e
+ S(4.8).

In short, the familiar pattern of recursive strong–weak alternations is common to both systems; the novel C15.P66
contribution of the weighted form is the di erentiation of the relative gaps between metrical levels. The

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
main musical e ect of this is to highlight the importance of the central tactus level(s). For present purposes,
weighting the contribution of each metrical level in this way enables us also to weight the relative strength
of the metrical dissonances a ecting them.

The ongoing relevance of the discrete model bears emphasizing because the shared method enables us to C15.P67
have the bene ts of weighted-values while still making the generalisations outlined in ‘Summary of the
Discrete-weighting Model’. Indeed, all of those generalizations still hold for the salience-weighted versions
of displacement dissonance of binary metres, with the exception of the di erences between dissonance
levels, for which there is now a narrow band of exception.The online supplements help to clarify this:
Supplement 1 www.oup.com/us/ohtm formalises and simpli es those claims for the general case (in both
discrete and salience-weighted conditions), and Supplement 2 www.oup.com/us/ohtm is an html
resource that enables free, interactive exploration of all possible pairings of common practice metres.

Figure 15.6

A stem chart for metrical weight, taking the example of two bars of 4/4 metre at crotchet = 100 bpm (IOI = 0.6 seconds) C15.F6
(reproduced from Gotham (2015a).

p. 293 Position- and Context-Specificity C15.S13

Even with the tempo-dependent system for weighting pulse-salience values, metrical positions still belong C15.P68
to classes of equivalent weighting, and so displacement dissonance is essentially measured on the basis of
movement from one metrical level to another, rather than to speci c positions. The strict equivalence of
metrical positions by level has been shown to be a useful formalization, but is also limited in capturing
details of the metre.

Were individual weightings for metrical positions to be introduced, the arithmetic mean method of C15.P69
assessing dissonance values would still generate comparative values, but those values would no longer fall
into neat, equivalent categories. Fewer generalizations are possible, but there do still seem to be identi able
trends. For instance, empirical studies of metrical position usage such as Prince and Schmuckler (2014)
show that position usage corresponds strongly to theoretical assertions of clearly graduated levels in
metrical structure, but they also suggest other e ects within levels such as a greater usage of positions late
in the cycle than those at the start.

The e ect that this would have on overall dissonant values de es simple summary, but we can characterize C15.P70
something of the di erent distributions of dissonance across metrical cycles. For instance, for D1 of our
example binary metre of cycle length 16, most of those comparisons formerly belonging to a single category

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
would remain relatively similar to one another as each level follows the same gradual increase. The
exception falls at the start/end of the cycle, where one value moves from being at the end (which attracts the
maximum within-level weighting) to being at the beginning (the minimum).

By contrast, for approximately half-cycle displacements of the same metre like D7, the corresponding C15.P71
di erences would be more evenly spread out. In short, D1 would emphasize larger changes in a concentrated
area around the cycle boundary with little di erence at other positions, while D7 would have less
individually striking di erences, though more positions would be more greatly di erentiated.

Among other limitations of the model at this stage, the context-blindness is problematic; future work could C15.P72
perhaps combine this model with formalizations of accent in musical contexts such as Volk’s computation
on the basis of actual onset usage in context (2008) and on the basis of actual onset usage in context and
Roeder’s formalization of accent across multiple parameters (2003).

Mixed Metres: Metrical Weighting C15.S14

18
Could this system be extended to mixed metres (5s, 7s…)? The standard Western music theoretic account C15.P73
of mixed metre sees the well-formedness constraints that admit only duple and triple groupings between
p. 294 levels extend to admit them also within levels. Thus, we add to the simple metres with cycles of length 2,
4, 6, 8, 12…‘new’ metres of length 5, for instance, but only in 2 + 3 or 3 + 2 groupings, not 4 + 1 or 1 + 4.

There is some psychological and music theoretic evidence to support this position, though it is not yet clear C15.P74
that handling these metres with the same kinds of hierarchical structures as simple metres is cognitively
19
appropriate, and it is doubly speculative to apply the new weighting systems outlined above. Those caveats
notwithstanding, we can certainly explore the rami cations that this system would have for metrical
dissonances, and perhaps even contribute something to the resolution of those issues in doing so.

Let us take the example of 7/8 in the 2 + 2 + 3 rotation. This would entail a discrete representation of: C15.P75

• 7 units = 1,1,1,1,1,1,1 C15.P76

• Grouped in 3 beats = 2,1,2,1,2,1,1 C15.P77

• Grouped into a bar (initial beat strongest) = 3,1,2,1,2,1,1 C15.P78

In addition to the problems with the discrete account already discussed, mixed metres pose the additional C15.P79
consideration of non-equivalent durations within the same metrical level. Here again, tempo/absolute
duration enables a possible response.

According to the de nition of this 7/8 metre set out above (with one triple and two duple beats), the ‘1’ level C15.P80
refers to a regular unit pulse; the ‘3’ level refers to a (potentially) regular 7-unit (bar) pulse; but on the ‘2’
level there are two di erent spans represented: both duple (2-unit) and triple (3-unit) beats. To re ect this
account of the metrical hierarchy, the two spans must both be represented. So, in addition to weighting each
pulse-level, we also need two distinct values in place of the second level, with values corresponding to S(2x)
and S(3x).
Figure 15.7

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Stem chart for 7/8 with crotchet = 100 (IOI = 0.6), giving two bars within the metrical window. Le : in the rotation 2 + 2 + 3; right: C15.F7
in the rotation 3 + 2 + 2.

As always, these values are tempo-dependent, so we turn to a tempo-metrical example for illustration. C15.P81
Figure 15.7 gives two versions of the metrical stem chart for two bars of 7/8: in the 2 + 2 + 3 arrangement
discussed so far on the left, and in an alternative 3 + 2 + 2 con guration on the right. The tempo for these
p. 295 examples corresponds as directly as possible to the binary example in Figure 15.6, with the 2-unit
crotchet beat at MM100. The values for these metres are also given in Table 15.4.

Table 15.4 Metrical weighting values for 7/8 at crotchet = 100 bpm, in two di erent rotations: 2 + 2 + 3 and 3 + 2 + 2. S stands for C15.T4
the individual pulse salience, and M for the combined metrical weightings

7/8 at C = 100 x S Pulse M

1 0.15 0.133486 Semi-quaver 0.133486

2 0.3 0.604448 Quaver 0.737934

4 0.6 1 Crotchet 1.737934

6 0.9 0.841754 Dotted-crotchet 1.713175

For 223

14 2.1 0.193109 Bar 1.931044

28 4.2 0.018916 2 × Bar 1.949960

For 322

14 2.1 0.193109 Bar 1.906284

28 4.2 0.018916 2 × Bar 1.925200

According to the model, the crotchet level is optimized at this tempo (100 bpm). Therefore, the dotted- C15.P82
crotchet beat corresponds to a non-optimal span and receives lower overall combined salience values than
20
its binary (non-dotted) equivalent. See the highlighted values in Table 15.4, and compare positions x = 1.2
and 3.3 (triple beats) with the x = 0.6 and 2.7 (duple) on the graph for the 2 + 2 + 3 rotation.

This leads to an overall weighting di erence between the 2 + 2 + 3 and 3 + 2 + 2 rotations. In the 3 + 2 + 2 C15.P83
rotation, this dotted beat falls at the start of the bar, leading to lower values at the highest levels for this
21
metre as compared with 2 + 2 + 3 (see the ‘Bar’ and ‘2 × Bar’ values on Table 15.4). This suggests that there
may be an in uence of tempo-dependent metrical weighting on rotational stability. At this tempo, the 3 + 2 + 2
rotation achieves a lesser overall metrical weighting; 2 + 2 + 3 may be considered the rotation better
optimized for metrical salience.

It is important to stress that these speci c outcomes are all tempo-dependent. At this tempo, the undotted C15.P84
value receives a greater weighting than the dotted version; however, for the kinds of fast tempi often used
for mixed metres, the dotted, longer beat length will be closer to the maximal salience range and thus
attract a greater metrical weight in this system.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
As we’re taking on metrical rotation, then we ought to address the preference of 2 + 2 + 3 and 3 + 2 + 2 over C15.P85
the remaining rotation, 2 + 3 + 2 (apparently across tempi and musical styles). One way to re ect this
priority of 2 + 2 + 3 and 3 + 2 + 2 over 2 + 3 + 2 would be to admit a ‘half-bar’ style weighting as part of the
p. 296 account of 7/8 where possible which accepts 4 + 3 arrangements but not 5 + 2, perhaps on the basis of
22
maximal evenness. This would give an additional 4 + 3 level for 2 + 2 + 3, and 3 + 4 level for 3 + 2 + 2, thus
supporting those preferred metres over 2 + 3 + 2, which would be excluded. This would require a relaxation
of the well-formedness criteria already discussed which require 2 or 3 groupings only at each level; we now
have groups representing 2 and 1 elements respectively—the 3-unit beat has not been grouped at this new
level. These alternative structural theories set out one way in which modelling of the kind outlined here may
help frame future research seeking to engage with these cognitive questions.

Mixed Metres: Metrical Dissonance C15.S15

If this kind of metrical weighting is indeed appropriate for mixed metres, then it allows us to include mixed C15.P86
metres in the method for quantifying metrical dissonance advanced above.

Due to the use of unequal spans within the same level, fewer categorical equivalences emerge between C15.P87
displacement types. With a 7-unit metre, for instance: it continues to be the case that D+ 1 = D+ 6 because D+ 6
= D−1, but D+ 1≠ D+ 3 (in any rotation), and so we cannot extend those generalizations of the ‘Summary of the
Discrete-Weighting Model’ this far.

While we may not be able to categorize types of mixed metrical displacement as easily as for the simpler C15.P88
common practice metres, we can still make comparisons using the overall average-per-position value. This
is speci cally designed to allow relative comparison of any dissonance type at any tempo. We might
therefore ask how displacements of a mixed metre against itself compare with such displacements of
common practice metres.

Continuing with the examples we have already encountered, the 7/8 (2 + 2 + 3) metre shown on the left side C15.P89
of Figure 15.7 displaced by a quaver (D+ 2) gives a weighted metrical dissonance value of 0.454 per unit
position. This relates directly to the binary metre for the same tempo (crotchet = 100) shown in Figure 15.6.
The same, quaver unit displacement (D+ 2) in that case gives 0.668, a considerably higher value. This suggests
that binary metres may be more adversely a ected by displacement dissonances than the most directly comparable
mixed metres. Perhaps the more stable a metre is (the higher its overall metrical weighting) the more it
stands to lose.

Grouping Dissonances and Cycle Lengths C15.S16

This comparison of metrical dissonance values across dissimilar metres is important not least because it C15.P90
enables an account for the other type of metrical dissonance: grouping dissonance. The method of averaging
per position in a cycle ensures that this system applies just as readily to grouping dissonances as to the
displacement dissonances which have been the focus of this study. There is not space for a full exposition of
grouping dissonances here, but some important comments are in order and Online Supplement 2
www.oup.com/us/ohtm allows free exploration.
p. 297 As displacement dissonance features two (or more) di erent displacements but of the same (one) metre, the C15.P91
‘whole cycle’ is necessarily equal in length to that shared metre. In grouping dissonance, by contrast, the
metres may be of di erent lengths, and so the cycle length of the dissonant combination must be taken
from the Lowest Common Multiple (LCM) of the two metres. For instance, the cycle length of a 16-unit
metre in any displacement is 16 units, whereas the cycle for that metre in a grouping dissonance with a 12-
unit metre is 48: LCM(16, 12) = 48. Clearly, cycle-lengths for grouping dissonance will tend to be much
longer than those for displacements dissonances, but whatever the cycle lengths, the arithmetic mean

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
allows us to compare in a general sense any kind of metrical dissonance, both grouping and displacement,
23
whatever the di erent cycle lengths involved. Online Supplement 2 www.oup.com/us/ohtm provides an
interactive guide for testing out these relative values, very much including grouping dissonance.

Perhaps the most basic metrical considerations that this account of grouping dissonance helps to illustrate C15.P92
is that which holds between two metres where the larger can be fully expressed by successive instances of
24
the smaller. For example, compare the ‘2/4’ and ‘2/2’ versions of the Bach example discussed—the latter
including one extra metrical level. The ‘grouping dissonance’ between those two metres only involves the
highest level—a small value, especially in the weighted form of this model for any realistic tempo.

Table 15.5 Discrete values for three hemiola-related grouping dissonances C15.T5

6∕8 vs 3∕4

6/8 3 1 1 2 1 1

3/4 3 1 2 1 2 1 Sum Mean

Di erence 0 0 1 1 1 0 3 0.5

2∕4 vs 3∕4

2/4 3 1 2 1 3 1 2 1

3/4 3 1 2 1 2 1 3 1

Di erence 0 0 0 0 1 0 1 0

6∕8 vs 2∕4

6/8 3 1 1 2 1 1 3 1

2/4 3 1 2 1 3 1 2 1

Di erence 0 0 1 1 2 0 1 0

More realistic grouping dissonances involve di erent metrical structures like 3/4 versus 6/8. Table 15.5 sets C15.P93
out discrete values for three hemiola-related metrical dissonances, somewhat analogous to those provided
for displacement dissonance at the start of this chapter. Again, these values provide a starting-point,
though there are signi cant issues associated with the discrete account. Note that the dissonance between
p. 298 2/4 and 3/4, for instance, is equivalent to that between 6/8 and 3/4 (i.e. a hemiola), simply with an
additional, lower level. That additional level sees the discrete system return a very di erent overall
dissonance value; the weighted model, by contrast, will only change signi cantly if the tempo and salience
of the added level warrant such a change.

Apart from questions of space, the lack of attention to grouping dissonance in the mathematical supplement C15.P94
re ects the fact that algebraic simpli cations are less forthcoming for grouping dissonances. Because the
two metrical structures are di erent, we no longer have equivalent values for corresponding levels. The
values can still be worked out easily enough (as Online Supplement 2 www.oup.com/us/ohtm shows), but
they do not make for simple equations of the kind provided for the displacement dissonances.

Conclusion C15.S17

Theoretic accounts of musical processes do well to incorporate the psychology of listening as part of C15.P95

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
building holistic and relevant models, even though that necessarily involves speculative operational
assumptions and generalizations. The model for quantifying metrical dissonance outlined in this chapter
has attempted to negotiate that balance with a concomitant trade-o between general usability and context
sensitivity. Perhaps the strongest aspect of the model is its synthesis of a broad range of core elements: both
metrical dissonance types, metres with any number of levels, and a range of relevant music-theoretic and
cognitive scienti c principles. This leads to some intriguing implications for the relative strength of
di erent metrical dissonances.

First, beyond merely reinforcing the relatively familiar music-theoretic notion that dissonance intensi es as C15.P96
one proceeds from the highest to the lowest levels, the model extends this to suggest the (unfamiliar) idea that
the di erences between those successive values also increase at each step. The mathematics for this identify a
lower limit to this latter correlation, though in practice the principle holds for almost the entire ‘window’ of
metrically useful pulse values (see Online Supplement 1 www.oup.com/us/ohtm ).

The model’s unusual breadth even extends to mixed metres. Again, this necessarily involves considerable C15.P97
speculation, but the exploratory approach taken here is rewarded with compelling principles that certainly
warrant further exploration and testing. First, in comparing the rotations of mixed metres, the model
suggests a possible formalization for the in uence of tempo-dependant metrical weighting on rotational stability
that may help to explain the widely attested preference for some rotations (2 + 2 + 3 and 3 + 3 + 2) over
others (2 + 3 + 2). Moreover, by developing a system that will compare any dissonance type between any
metrical structure, we are able to pursue comparisons between mixed and non-mixed metres. Here, the
model suggests that the simple, binary metres appear to be more adversely a ected by displacement dissonances
than the most comparable mixed metres. This accords with the intuition that already ‘unstable’ mixed metres
may admit metrical changes more readily than ‘stable’ common-practice metres.

These are intriguing principles that will, it is hoped, help motivate and guide further study, but they are only C15.P98
p. 299 preliminary steps on the long road to developing a rich and robust account of how metrical dissonance
‘really’ works. This model makes only preliminary inroads into the complex psychology of metrical
listening, and gives almost no account of the highly varied metrical practices across the extraordinarily
diverse repertoires for which this model might potentially be relevant. The main text raises some speci c
issues of particular relevance to advancing the theory as it stands. To this could be added many further
considerations, as there is a wide range of repertoire contexts for which some kind of hierarchical, metrical
approach is pertinent, but for which the speci c constraints of 2- and/or 3-grouping of nominally regular
pulses would not be appropriate.

As so often, especially at the outset of a developing theory, there are di erent approaches to pursue here, C15.P99
and the biggest leap forward may be accomplished by theories that are able to synthesises the ‘best bits’ of
each. An alternative method for quantifying metric dissonance recently proposed by Reale (2019) charts a
di erent course that remains conceptually closer to Krebs in its focus on note onsets and the distances
between them, operationally de ning ‘a quantity of metric dissonance’ as the ‘durational distance between
two associated metrical onsets’ (Reale 2019: 149). In short, where the present study measures di erence
between items at each point in time, Reale measures the time that elapses between ‘equivalent’ events.
There are bene ts and detractions from each approach. Reale’s method arguably does a better job of
accounting for the rhythmic aspect of note-by-note events in an analytical context, (especially for
displacement dissonance), while the model presented here perhaps accounts more e ectively for metrical
levels and cognitive matters.

While there are many possible avenues for engaging with open research questions on the theoretical side, C15.P100
there is perhaps enough already here to meaningfully support scholars and practitioners working on pieces
that involve metrical dissonance. A great deal of musical thought depends on notions of relative similarity

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
and di erence, including, for instance, even commonplace distinctions between ‘close’ and ‘distant’ keys.
Correspondingly, much systematic and computational work is dedicated to formalizing these notions in
support of musical analysis. We do this not to close the proverbial book on the matter at hand—no one could
credibly claim to have provided the ‘de nitive’ solution for how to map proximity and distance in musical
keys, for instance. Instead, we aim to get enough of a grasp on any particular device in a given context to
develop our understanding of the music in question. Insofar as this model helps to advance the state of play
for understanding metrical dissonance, it may serve to reinvigorate our understanding of repertoires for
which metrical dissonance is an important factor, and in so doing, perhaps broaden the range of repertoires
we feel prepared to discuss and perform. Scholars and practitioners cannot wait for ‘perfect’ theory any
more than music theory can wait for ‘conclusive’ cognitive science. We move forwards and outwards,
together, incrementally.

Notes
1. The grounds for a correspondence between the overtone series and the tonal system, however, are fraught with problems, C15.N1
not least of which being the unbalanced status of the ʻsub-dominantʼ as important but ill-fitting.

p. 300 2. Note, of course, that we are dealing with a logarithmic scale for pitch, rather than a linear one. C15.N2

3. This include the notion of dissonance by ʻproximity to consonanceʼ (Krebs 1999: 57) C15.N3

4. The equivalence of the two is perhaps overstated by the iconic ʻcontinuumʼ propounded by Cowell (1930) and C15.N4
Stockhausen (1957). See London (2002) for ʻsome non-isomorphismsʼ between the two parameters.

5. Each level is represented by a succession of dots, Kramer (1988) sets out the same idea with dashes, and Cooper and C15.N5
Meyer (1960) invoke a parallel with linguistic ʻmetrical feetʼ. See London (2012: 78/202 n. 2) for a summary.

6. Krebs (1999) uses the term ʻsubliminal dissonanceʼ for cases where all musical parameters unanimously contradict the C15.N6
notated metre.

7. Apart from Krebs (1999), see Cohn (1992b; 1992a; 2001); McClelland (2006); Malin (2008); Mirka (2009); Malin (2010); C15.N7
McClelland (2010); Murphy (2009); Murphy (2012); Willner (2013).

8. See esp. Biamonte (2014) for an introduction. C15.N8

9. See e.g. Lewin (1981); Cohn (2001); London (2012); and for the rhythmic case, Yust (2020). C15.N9

10. Reale (2019) is a notable exception, likewise motivated by the analytical need to assess ʻwhich of these events is more C15.N10
metrically dissonant?ʼ and the understanding that ʻit would be useful to have a formalized method for measuring degrees
of dissonance, allowing for the quick comparison of diverse metric eventsʼ (p. 147). Reale and I share a common goal, but
diverge in our approaches, as will be discussed in the concluding, outlook section.

11. See Cohn (2016) for some more examples and a recent theorization of the structure. C15.N11

12. See Polak (2010: nn. 9 and 10) for further references rejecting polymeter, particularly in the context of Jembe music. C15.N12

13. See Huron (2016: ch. 5) for a summary of the perceptual evidence in support of the ʻtrill thresholdʼ. C15.N13

14. E.g. a geometric mean could not cope with the numerous zero values for positions of equivalent metrical weighting, C15.N14
particularly for the lowest metrical levels. Alternatively, the Hamming distance could provide a di erence metric for the
two strings, but as it insists on binary values (0/1)—the mere presence or absence of di erence—it throws out the metrical
weighting values. The Hamming distance is also one of many metrics (including Euclidean distance and cosine similarity)
that do not lend themselves to the comparison of cases with di erent cycle length.

15. In Lerdahl and Jackendo (1983)ʼs terms, higher levels are ʻopen to interpretationʼ (p. 22). C15.N15

16. For some (Marissen 1990), this ʻramblingʼ and ʻthematically homogenousʼ texture may not be a desirable quality. See C15.N16

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Colahan (2008) for further metric analyses of this movement alongside the rest of the Brandenburg concerto cycle.

17. For an discussion of the kind of tempo Bach might have had in mind for his alla breve works, see Jerold (2014). The e ect C15.N17
of tempo on this model of metrical dissonance is discussed below.

18. For an alternative theorisation of metrical dissonance in mixed metres, see Smith (2018). C15.N18

19. I explore this question in Gotham (2017). C15.N19

20. The dotted-crotchet value (given by S(x) + S(2x) + S(6x) = 1.713) is lower than that for the optimized crotchet (S(x) + S(2x) + C15.N20
S(4x) = 1.738).

21. For 3 + 2 + 2, the bar-beginning metrical weight is given by S(x)+S(2x)+S(6x)+ S(14x) = 1.906, less than that for the 2 + 2 + 3 C15.N21
rotation, given by S(x) + S(2x) + S(4x) + S(14x) = 1.931.

22. See Clough and Douthett (1991). C15.N22

p. 301 23. E.g. the grouping dissonance of 3 (2 levels: 2–1–1) against 4 (3 levels: 3–1–2–1) = 9/12 = 0.75, while the displacement C15.N23
dissonance of that 4 against its anti-phase displacement is only 2/4 = 0.5.

24. In Gotham (2015b) I dub this the ʻSaturation relationʼ, but the concept appears in accounts of musical metre as diverse as C15.N24
Kirnberger (1982[1776]) and Longuet-Higgins and Lee (1984).
References C15.S18

Biamonte, N. (2014). Formal functions of metric dissonance in rock music. Music Theory Online 20(2). C15.P101
Google Scholar WorldCat

Clough, J., and Douthett, J. (1991). Maximally even sets. Journal of Music Theory 35(1/2): 93–173. C15.P102
http://www.jstor.org/stable/843811

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Cohn, R. (1992a). The dramatization of hypermetric conflicts in the scherzo of Beethovenʼs ninth symphony. 19th-Century Music C15.P103
15(3): 188–206. http://www.jstor.org/stable/746424
Google Scholar WorldCat

Cohn, R. (1992b). Metric and hypermetric dissonance in the menuetto of Mozartʼs Symphony in G Minor, K. 550. Intégral 6: 1–33. C15.P104
http://www.jstor.org/stable/40213939
Google Scholar WorldCat

Cohn, R. (2001). Complex hemiolas, ski-hill graphs and metric spaces. Music Analysis 20(3): 295–326. C15.P105
http://www.jstor.org/stable/854473
Google Scholar WorldCat

Cohn, R. (2016). A platonic model of funky rhythms. Music Theory Online 22(2). C15.P106
http://mtosmt.org/issues/mto.16.22.2/mto.16.22.2.cohn.html
Google Scholar WorldCat

Colahan, E. (2008). Metric conflict in the Brandenburg Concertos of J. S. Bach. Masterʼs thesis, University of Denver. C15.P107

Cooper, G., and Meyer, L. (1960). The rhythmic structure of music. University of Chicago Press. C15.P108
Google Scholar Google Preview WorldCat COPAC

Cowell, H. (1930). New musical resources. Knopf. C15.P109


Google Scholar Google Preview WorldCat COPAC

Gotham, M. (2015a). Attractor tempos for metrical structures. Journal of Mathematics and Music 9(1): 23–44. C15.P110
Google Scholar WorldCat

Gotham, M. (2015b). Meter metrics, Music Theory Online 21(2). C15.P111


http://www.mtosmt.org/issues/mto.15.21.2/mto.15.21.2.gotham.html
Google Scholar WorldCat

Gotham, M. (2017). Hierarchy and position usage in mixed metres. Journal of New Music Research 46(2). C15.P112
Google Scholar WorldCat

Huron, D. B. (2016). Voice leading: the science behind a musical art. MIT Press. C15.P113
Google Scholar Google Preview WorldCat COPAC

Jerold, B. (2014). Tempi in the era of Bach. Musical Times 151(1927): 85–96. C15.P114
Google Scholar WorldCat

Kirnberger, J. P. (1982[1776]). The art of strict musical composition [Die Kunst des reinen Satzes in der Musik] (trans. and ed. C15.P115
D. Beach and J. Thym). Yale University Press.

Kramer, J. D. (1988). The time of music: new meanings, new temporalities, new listening strategies. Schirmer. C15.P116
Google Scholar Google Preview WorldCat COPAC
Krebs, H. (1999). Fantasy pieces: metrical dissonance in the music of Robert Schumann. Oxford University Press. C15.P117
Google Scholar Google Preview WorldCat COPAC

Lerdahl, F., and Jackendo , R. (1983). A generative theory of tonal music. MIT Press. C15.P118
Google Scholar Google Preview WorldCat COPAC

Lewin, D. (1981). On harmony and meter in Brahmsʼs Op. 76, No. 8. 19th-Century Music 4(3): 261–265. C15.P119
http://www.jstor.org/stable/746699

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

London, J. (2002). Some non-isomorphisms between pitch and time. Journal of Music Theory 46(1/2): 127–151. C15.P120
http://www.jstor.org/stable/4147679
Google Scholar WorldCat

London, J. (2012). Hearing in time: psychological aspects of musical meter (2nd edn). Oxford University Press. C15.P121
Google Scholar Google Preview WorldCat COPAC

p. 302 London, J. (n.d.). Rhythm. In Oxford Music Online (http://www.oxfordmusiconline.com/)ʼ. Oxford University Press. C15.P122
Google Scholar Google Preview WorldCat COPAC

Longuet-Higgins, H. C., and Lee, C. S. (1984). The rhythmic interpretation of monophonic music. Music Perception: An C15.P123
Interdisciplinary Journal 1(4): 424–441. http://www.jstor.org/stable/40285271
Google Scholar WorldCat

Malin, Y. (2008). Metric analysis and the metaphor of energy: A way into selected songs by Wolf and Schoenberg. Music Theory C15.P124
Spectrum 30(1): 61–87. http://www.jstor.org/stable/10.1525/mts.2008.30.1.61
Google Scholar WorldCat

Malin, Y. (2010). Songs in motion: Rhythm and meter in the German Lied. Oxford University Press. C15.P125
Google Scholar Google Preview WorldCat COPAC

Marissen, M. (1990). Relationships between scoring and structure in the first movement of Bachʼs Sixth Brandenburg Concerto. C15.P126
Music and Letters 71(4): 494–504. http://www.jstor.org/stable/736819
Google Scholar WorldCat

McClelland, R. (2006). Metric dissonance in Brahmsʼs Piano Trio in C Minor, Op. 101. Intégral 20: 1–42. C15.P127
http://www.jstor.org/stable/40214027
Google Scholar WorldCat

McClelland, R. (2010). Brahms and the scherzo: Studies in musical narrative. Ashgate. C15.P128
Google Scholar Google Preview WorldCat COPAC

Mirka, D. (2009) Metric manipulations in Haydn and Mozart: Chamber music for strings, 1787–1791. Oxford University Press. C15.P129
Google Scholar Google Preview WorldCat COPAC

Murphy, S. (2009). Metric cubes in some music of Brahms. Journal of Music Theory 53(1): 1–56. C15.P130
http://www.jstor.org/stable/40606877
Google Scholar WorldCat

Murphy, S. (2012). Septimal time in an early finale of Haydn. Intégral 26: 91–121. http://www.jstor.org/stable/23629591 C15.P131
Google Scholar WorldCat

Parncutt, R. (1994). A perceptual model of pulse salience and metrical accent in musical rhythms. Music Perception 11(4): 409– C15.P132
464. http://www.jstor.org/stable/40285633
Google Scholar WorldCat
Polak, R. (2010). Rhythmic feel as meter: Non-isochronous beat subdivision in jembe music from Mali. Music Theory Online 16(4). C15.P133
Google Scholar WorldCat

Prince, J. B., and Schmuckler, M. A. (2014). The tonal-metric hierarchy: A corpus analysis. Music Perception 31(3): 254–270. C15.P134
http://www.jstor.org/stable/10.1525/mp.2014.31.3.254
Google Scholar WorldCat

Reale, S. (2019). The calculus of finite (metric) dissonances. Music Theory Spectrum 41(1): 146–171. C15.P135

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470064 by National Science & Technology Library user on 26 May 2023
https://doi.org/10.1093/mts/mty028
Google Scholar WorldCat

Roeder, J. (2003). Beat-class modulation in Steve Reichʼs music. Music Theory Spectrum 25(2): 275–304. C15.P136
http://www.jstor.org/stable/10.1525/mts.2003.25.2.275
Google Scholar WorldCat

Schoenberg, A. (1951). Style and idea. Williams & Norgate. C15.P137


Google Scholar Google Preview WorldCat COPAC

Smith, J. (2018). Metric dissonance in non-isochronous meters. PhD thesis, University of North Texas. C15.P138
Google Scholar Google Preview WorldCat COPAC

Stockhausen, K. (1957). … wie die Zeit vergeht …. Die Reihe 3: 13–42. C15.P139
Google Scholar WorldCat

van Noorden, L., and Moelants, D. (1999). Resonance in the perception of musical pulse. Journal of New Music Research 28(1): C15.P140
43–66.
Google Scholar WorldCat

Volk, A. (2008). Persistence and change: Local and global components of metre induction using inner metric analysis. Journal of C15.P141
Mathematics and Music 2(2): 99–115. http://www.tandfonline.com/doi/abs/10.1080/17459730802312399
Google Scholar WorldCat

Willner, C. (2013). Metrical displacement and metrically dissonant hemiolas. Journal of Music Theory 57(1): 87–118. C15.P142
Google Scholar WorldCat

Yust, J. (2020). ʻGeneralized Tonnetze and Zeitnetze, and the topology of music concepts. Journal of Mathematics and Music 14(2): C15.P143
170–203. https://doi.org/10.1080/17459737.2020.1725667
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
CHAPTER

16 Maelzel, the Metronome,


C16 and the Modern Mechanics of
Musical Time 
Alexander E. Bonus

https://doi.org/10.1093/oxfordhb/9780190947279.013.16 Pages 303–C16.P255


Published: 08 December 2021

Abstract
Johann Nepomuk Maelzel, despite being most recognized today for inventing the clockwork
metronome, was one of the most famous automata showmen of the nineteenth century. This chapter
begins by o ering a reception history of Maelzel, the metronome, and his automata, and exploring the
cultural signi cances underlying his clockwork creations across the Industrial Age. As numerous
accounts maintain, Maelzel’s automata projected decidedly inhuman performance practices. His
automata emblematized a machine culture that ran in direct opposition to the subjective ‘artistry’
championed by many skilled performers and composers over the century. This study subsequently
addresses the discord between Maelzel’s age and ours regarding the values of musical time and
performance practices: those metronomic qualities largely rejected by Maelzel’s musical
contemporaries are often vehemently endorsed today by many professional musicians and educators
who apply mechanically precise tempos and rhythms to all musical repertoires. This history ultimately
confronts the veiled ‘metronome mentality’ found throughout contemporary performance culture,
which neglects many musical-temporal aesthetics and rhythmic qualities from a pre-industrial, pre-
metronomic past.

Keywords: rhythm, metre, tempo, technology, performance, regulation, automatic, automaton,


Beethoven, precision
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

AUTOMATIC metronomes, as used in present-day musical activities, hold a seldom-articulated but often C16.P1
assumed function: to objectively measure the accurate tempo of a musical composition. In practice and
pedagogy, they enforce the supposedly correct rhythm of a movement, passage, or exercise. Attesting to its
own essentiality, the typical turn-of-the-century Maelzel-style metronome was accompanied by the
statement: ‘A Metronome is an instrument which indicates at which speed a piece of music is to be played’
(Bingham 2012: 33, 34). One of the rst U.S. patents for an electric metronome, led in 1902, summarizes
the technology’s applications, which, for modern musical practitioners, might seem self-evident:

In the study of music the [clockwork] metronome has come to be an important element not only in C16.P2
determining the tempo that a certain composition shall be played at, but in aiding the student in
technical exercises to maintain a precision of rhythm in any tempo from largo to presto.

(U.S. Patent No. 734,032, 1903)

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
When many re ect upon the fundamentals of musical time, they invariably think about metronomic C16.P3
(continuous beats per minute) measurements. For over a century, promotions of automatic metronomes
have bolstered the view that musical time and precise metronomic action are fundamentally uni ed in
theory and practice. How many current performers and pedagogues would nd these dictums from an 1897
advertisement contentious?

p. 304 The Metronome tells you the rate of movement at which you ought to play or sing a piece. C16.P4

The Metronome tells you whether you can go through the piece in unbroken measure. C16.P5

The Metronome is your rhythmic guide, philosopher and friend, without which at hand for C16.P6
consultation you will be all at sea.

The Metronome tells you whether you are playing your slow movements without hurrying on the C16.P7
one hand or without hanging back.

(Advertisement, 1897)

However de nitively stated, the true principles of musical time extend past commercial endorsements, both C16.P8
old and new. Beyond modern promotions—and the seemingly innocuous MM (Maelzel’s Metronome)
numbers lingering on old music editions—metronomes emblematize deeper societal values about
temporality, rhythmicality, and human action itself. The ancients well understood this fact: time is not a
mere, measurable quantity. Indeed, what patents and advertisements—and many modern music
instructions—fail to explain is that when we seek to quantify time, we are subscribing to a host of unspoken
beliefs in and about qualitative actions and movements. What modern individuals often consider the ‘right
time’ is an acculturated, technologically dependent value system that no clock can fully document.

A Preamble on Musical and Mechanical Precision C16.S1

Although this chapter cannot comprehensively address the mutable concepts of time throughout the course C16.P9
of civilization, it can be stated that new technologies bring new paradigms of time-knowing, and those new
technological paradigms are often employed to reorient, reorganize, and rede ne communal and individual
actions. Behaviours in time are entailed to the ways we choose to, or are taught to, measure time. It would be
a great challenge to arrive precisely ten minutes early to a business meeting, aided only by the sundial xed
in your neighbour’s backyard. It would be equally di cult to catch the 9.55 a.m. train with sole reference to
the pealing bells from a local monastery. For modern activities, we require more automatically precise,
personally accessible, and globally connective time-tellers.

The measurement of time, far from being an absolute, objective truth, is therefore an undeniable C16.P10
sociocultural construction—a collection of learned customs, ideals, and practices, which imparts
importance to one measurement construct above another. Take for example the modern stopwatch: what it
promotes, and what modern users value in it (whether they explicitly state it or not), is automatic precision
—a major theme that extends to modern metronome use. Automatic precision can be de ned as the
continuous and exact dividing of larger durations into equal, smaller units: seconds, milliseconds, and so
on.

In science and industry, precision measurements from the latest computerized time-tellers are important, C16.P11
p. 305 if not essential. Yet considering so many quotidian activities, precision is not everything—in many cases
it is meaningless. When batteries in a digital stopwatch expire, the measurement of milliseconds, alongside
the knowledge of milliseconds, disappears. (Who can accurately count a series of milliseconds unaided by

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
machines? The very thought is not intrinsic or intuitive to the human experience.) The temporal concept is
bounded to the technological construct of it. Yet the hourglass, unencumbered by the arti ciality of an
electronic millisecond-readout, remains capable of de ning an overall duration for cooking, studying,
reading, practising, or all other activities for which automatic precisions are irrelevant. Hourglasses, while
relatively imprecise technologies, are appropriately accurate time-tellers for many common purposes.

The corollary: musical time could be—and once was—measured through more natural, non-automatic C16.P12
paradigms of motion, which did not require modern digital metronomes and the subdivided beats per
minute (bpm) entailed by them. Historically non-mechanistic (pre-metronomic) enactments of the pulse,
of metre, and of rhythm projected di erent modes of musical movement. Disregarding the electronic
clickers so relied upon today, a hand gesture, a walker’s stride, and a pendulum swing can each measure
‘tempo’—but the kind and quality of that measurement, hence the fundamental de nition of ‘tempo’,
varies from technique to technique. As I have argued elsewhere, these di ering qualities of time-reference
should be a primary concern when nding the ‘right’ movements for past and present musical
compositions. The fact should be reiterated: musical time was a quality of rhythmic motion before it ever
became a precise quantum of clockwork sound (Bonus 2013; 2014; 2017). Musical time existed before
metronomes. Prior to the prevalent use of beat-machines, ‘tempo’ signi ed far more than the reductive
bpm quanti cation of it.

So how did modern musical culture come to value clockwork-oriented tempo constructs, in which C16.P13
automatic machines signal ‘the rate of movement at which you ought to play or sing a piece’? How did such
an aesthetic imperative—the often-unquestioned requirement to practise and perform under the guidance
of clockwork controls—prevail in pedagogical and professional arenas? Under what unacknowledged
motivations did modern culture’s sense of musical time become linked to a metronomic modality of action,
to such an extent that someone as eminent as Herbert von Karajan (1908–9) proclaimed in 1989:

But [my brain] is not a computer. I trained it with [automatic] metronomes. And I still test myself. I C16.P14
can walk in 120 [sonic bpm] and sing in 108; and if you ask me to sing in 105 now, I will manage it. If
I get it wrong, I feel it with my whole body. And in the orchestra, if a solo comes in slower or faster
[in sonic-metronomic time], I sense it right away; it makes me feel uneasy.

(Osborne 1989: 97)

Given Karajan’s late-life stance regarding metronomic tempo reproduction, it seems likely that other C16.P15
modern professionals and students, so obsessed with mechanical beat precision, fail to consider how tempo
clocks came to faithfully dictate musical movement in the rst place.

p. 306 This chapter therefore o ers a broad diachronic survey of changing sociocultural relationships to the C16.P16
metronome and the precise, automatic quality of time it demarcates. This critical history traces the
emergence and eventual acceptance of the metronomic reference of musical movement, while addressing
the technology’s current impacts upon modern musical culture. Through a host of primary-source
evidence, it is argued that the precision-oriented metronomic measurement of tempo which has emerged
over the last two centuries has helped to rede ne and reorient the rhythmical ideals of modern musicians at
large.
Discovering the relevance of metronomic tempo within the tenets of nineteenth-century musicality means C16.P17
re ecting upon the individual most credited for inventing the technological framework: Johann Maelzel
(1772–1838). In order to gain greater historical-cultural perspective on the once-radical metronomic
meaning of musical time, this survey rst reconsiders the machine’s earliest, most dogged promoter,
alongside his picaresque career and his other clockwork devices.

Following from Maelzel’s biography, a nineteenth-century reception history highlights prominent voices C16.P18
critical of the very notion of mechanical tempo reference. As will be shown, the clockwork metronome

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
brought with it new, radical qualities and concepts of musical temporality. And those new mechanical
paradigms provoked concerns, contentions, and dire criticisms regarding the very nature of living
performance. Indeed, while business interests progressively promoted tempo-technologies to the lay
public, eminent musicians often heard artistic falsity, damage, and destruction in this newly applied
clockwork contrivance. An automatic machine projected the inverse of artistry—the antithesis of temporal
expression and human-engendered musical motion. They claimed that the metronomic sound-beat was
never, and could never be, the a priori indication of musical tempo. Historical musical movement in time
was not so easily told with the mere purchase of a tempo clock.

Evidence unfurls to suggest that, although Maelzel’s clockwork technology did not instantaneously usurp C16.P19
traditional rhythmic epistemologies and performance practices, many modern conceptions of musical time
eventually became intertwined with clockwork references and regulations. Indeed, in a modern world
replete with automatic metronomes and their tempo measurements, many performers have grown
accustomed to believing in a ‘tempo’, rst and foremost derived from automatic sound-markers. The nal
section of this chapter therefore details what I have previously dubbed this ‘metronomic turn’, whereby a
once-novel technology became reassessed as a normal, necessary component for twentieth-century
pedagogies and practices.

The trend is still evident through contemporary musical endeavors: metronomic sound and action, as a C16.P20
general concept, has in many cases overtaken past, prevailing values and practices of musical movement. As
the conclusion to this survey suggests, musical time—as an aesthetic of modern musicality—is, in many
cases, entailed to technological frameworks. For numerous musical pedagogues and practitioners today,
musical time is fundamentally what automatic metronomes dictate it to be.

Ultimately this brief sociocultural history of automatic metronomes, which encapsulates their origins, C16.P21
p. 307 changing uses, and intended purposes, goes well beyond the functioning state of Beethoven’s machine or
his supposedly precise, authorial intentions. This metronome history is more concerned with modern
tendencies and current sociocultural relationships to these automatic technologies. It challenges us to
reconsider metronome users of the modern age, what they seek from tempo machines, and
correspondingly, what they might falsely presume about historical composers’ intentions.

As will be shown, nineteenth-century performing artists and composers seldom revered modernist C16.P22
precision-time constructs, or scienti cally imbued training methods, or metronome-dependent processes
for shaping performance productions. The chief aim of this present chapter is to trace the entire enterprise
of the metronomic temporal-construct—and how automatic beat-technology became such a signi cant
factor in shaping the theories and practices of modern musicality.
Recounting the Professor of Music and Mechanics C16.S2

Although Maelzel’s life and legacy in toto has seldom been addressed by modern musical scholarship, those C16.P23
concerned with the topic of musical temporality ought to know the profound historical impacts of Maelzel’s
technoculture. The following biographical sketch suggests that Maelzel’s most signi cant historical
contribution was not simply the ‘invention’ of the clockwork metronome. Rather it was the presentation of
a new and unusual clockwork culture to a rapidly industrializing civilization. For his nineteenth-century

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
audiences, Maelzel rst displayed the promise, possibilities, and pitfalls of a mechanized world—one
populated with self-moving motors and motorized activities. It is a world that today seems both
commonplace and somewhat unremarkable. Before the globe was organized through time zones, before
coal-burning factories darkened city skies, Maelzel was the rst popular gure to endorse a cult of
automatic machinery and, with it, automatic performance activity. ‘The Professor of Music and Mechanics’,
as Maelzel was announced to America (Fiske 1859: 428), was the premier purveyor of mechanically self-
moving performances. Automating musical behaviour was Johann Maelzel’s professional speciality.

Maelzel moved to North America from Europe in 1825 to continue his career as the premier exhibitor of self- C16.P24
moving machines, or automata. Maelzel o ered the growing nation visions of mechanical wonder, mystery,
and ingenuity, as he produced his automata shows in cities across America. An 1837 article praises Maelzel’s
mechanical menagerie, which included ‘some of the most beautiful and splendid of automatic machines’,
such as ‘automaton speaking gures […] the Panharmonicon […] the automaton trumpeter […] equestrian
automata […] slack rope dancers […] the melodium […] the automaton charlatan […] and “last not least”,
that unique and most masterly combination of music, mechanism and design, the grand and appalling
p. 308 panoramic spectacle of the con agration of Moscow’ (Maelzel 1837). The clockwork metronome
frequently represented one meagre entry in Maelzel’s larger oeuvre of self-moving machines. Henry Edward
Krehbiel noted that Maelzel ‘as [he had] in Europe […] depended for a livelihood on exhibitions of his
mechanical contrivances’ (1898: 233). Indeed, throughout the century, Maelzel was often better known for
his mechanical-musical displays than for his eponymous tempo clocks (Berhard 1828: 197–198).

The archetypal salesman, Maelzel was also unsurpassed in the art of public persuasion. Through C16.P25
entrepreneurialism, ambition, and self-promotional cunning, Maelzel built a worldwide reputation as
mechanical mastermind. Some recollected the showman’s charm and warmth during exhibitions (Grund
1837: 78). ‘The mechanician was not only a man of unquestionable inventive genius, but he also understood
the public’, wrote Beethoven scholar Alexander Wheelock Thayer, adding that Maelzel ‘knew as by instinct
how to excite and gratify curiosity without disappointing expectation, and had the tact and skill so to
arrange his [automata] exhibitions as to dismiss his visitors grateful for an amusement for which they had
paid’ (Thayer 1921: 252). Circus impresario P. T. Barnum recounted the encouragement he received from
Maelzel, a shrewd businessman whom Barnum (1855: 58) considered, ‘the great father of caterers for public
amusement’.

Yet these nineteenth-century portraits of Maelzel as hero-inventor are muddled by many who recalled his C16.P26
superior talents in humbuggery. In 1857, American scientist John Dela eld invoked Maelzel’s name with the
vast history of public deception (1857: 20). ‘Everybody talks sense now-a-days. But how many are there
who can talk successful folly and gain the reputation of wisdom by it?’ re ected the American lawyer
William Wirt, adding, ‘That is a species of mental legerdemain which puts a man on a level with the far-
famed Maelzel, the exhibitor of the Androides, compared with whom the Chief-Justice [John Marshall]
himself is but an everyday sort of man’ (Kennedy 1850: 201). For Wirt, the greatest legal and logistical minds
of the day could not match Maelzel’s ability to sell mechanical claptrap to the masses.

Maelzel’s supreme skills at entertaining and misdirecting audiences were seldom questioned, yet many C16.P27
mysteries remain concerning his actual mechanical expertise. Maelzel often adopted, copied, or bought
many of the renowned automata attached to his name. He purchased his most perplexing automaton, the
chess player known as ‘the Turk’, from renowned Habsburg engineer Wolfgang von Kempelen (1734–1804),
after the old inventor found the android too taxing to display (Metzner 1998: 183). It is well documented
that Johann co-opted Dietrich Winkel’s double-pendulum technology, which became the operative tempo-
keeping component in the eponymous Maelzel’s Metronome (MM). The 1884 Encyclopedia Britannica
labelled Maelzel an ‘imposter’ and pirate for his sordid capitalization on Winkel’s creativity (p. 207). The
Grove Dictionary of Music and Musicians (1889) posited that ‘Maelzel was evidently a sharp, shrewd, clever
man of business, with a strong propensity to use the ideas of others for his own bene t’ (vol. 2: 194–195),

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
while C. G. Hamilton (1916) con rmed that the entertainer ‘possessed a remarkable combination of
inventive genius, business ability, and readiness to appropriate unscrupulously to his own use the products
p. 309 of others’ brains’ (p. 5). Johann Maelzel’s long-belated obituary in the Athenaeum (appearing in 1855, the
year of his brother’s death) harshly concluded:

As a man, Maelzel seems to have been quarrelsome, extravagant, and unscrupulous. He can only be C16.P28
ranked amongst those empirics whose cleverness almost amounts to genius. Had he possessed a
larger amount of culture and of conscience, he might have done service to high Art.

(Timbs 1856: 94)

Some historians doubt whether Maelzel invented any machines at all. Organologist Alexander Buchner C16.P29
suggested that his brother Leonard devised many of the early automata that Johann exhibited throughout
Europe. ‘The brothers are therefore very often confused even in specialist works on the subject’, Buchner
remarks, adding, ‘It is in fact di cult to say exactly where the work of one ends and the other begins’
(Buchner n.d.: 79–80, plates 49–52). For this reason, some nineteenth-century reports, biographies, and
obituaries mistakenly interchange their rst names.

Maelzel’s public image was nothing if not protean. His multifaceted legacy combined the engineer, C16.P30
inventor, magician, salesman, and charlatan in ever-shifting, ever-con icting measures.

As technological showman, Maelzel drew both praise and scorn throughout the nineteenth century. Critics C16.P31
noted some dire outcomes arising from Maelzel’s mechanical performances. Some were startled that
machines could begin to pattern, however imperfectly, the complex behaviours of living, thinking beings
(Holmes 1863). As we will continue to see, others contemplated the alternative potential for human agency
to sublimate to the processes of automatic technology. The relationship between Maelzel’s Turk and its true
activating force—a constrained human chess player (Figure 16.1)—emblematizes Maelzel’s technocultural
legacy all too clearly.

In 1859 the Turk remained ‘a piece of mechanism historically more curious than any other the world has C16.P32
ever seen’ (Fiske 1859: 483). Its eerie legend grew through chess histories, ctions (Ho mann 1908),
numerous journal articles (the most famous of which was penned by a young Edgar Allen Poe (1836)), and a
memoir about Napoleon (Constant 1907: 1895). The Turk’s uncanny behaviour warranted this sustained
fascination. During a match, it moved one arm while its cold eyes scanned left and right. Upon achieving the
winning position, it said, ‘Check’. The Turk was also impetuous, prone to ts of physical violence, as well as
moments of deep re ection. But it made odd sounds, and as the match progressed, Maelzel would open the
Turk’s integral desk and tinker with its inner workings. Few recognized the great magician’s powers of
misdirection at work.

The Turk’s human capacities for creativity and intelligence were mystifying. It begged the question: could a C16.P33
self-moving automaton with a sensible mind be considered an automaton at all? The existential conundrum
drives Hanna F. Gould’s 72-line homage, ‘To the Automaton Chess Player’ (1829): ‘Thou wond’rous cause
of speculation— / Of deep research and cogitation, / Of many a head, and many a nation— / While all in vain
p. 310 / Have tried their wits to answer whether / In silver, gold, steel, silk, or leather, / Or human parts, or all
together, / Consists thy brain!’

Figure 16.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Maelzelʼs Automaton Turk was famed throughout the century as a clockwork chess phenomenon. Yet the pseudo-automaton C16.F1
was motivated by a hidden human performer. Cleverly confining the living player within the androidʼs mechanical façade made
the magic trick work (Hopkins 1897: 371; Arrington and Ohl 1960: 60).

Maelzel’s Turk was an outlier-automaton in nearly all respects: most self-moving machines were not C16.P34
revered for their seemingly human wits or intelligence. Clockwork action was once considered the antithesis
of wilful behaviour (Descartes 1850: 97). Indeed, an automaton was a prime metaphor for the supremely
mindless, rote-acting dolt. A being unmoved by nature or beauty; a hollow and expressionless gure
(Minshull 1803: 34), the automaton was the puppet of its inventor (Holcroft 1798: 233). For playwright
Thomas Holcroft (1745–1809), an imbecilic character was ‘a thing—who, say the best, was but an idiot, an
automaton’ (1839: 375). These tropes of automatic idiocy were rooted in technological realities. Automata
were limited and constricted by the arti cial technology that facilitated their self-movement. Gears,
springs, cogs, and levers could drive a device without constant human intervention, but motorized
components did not imbue an automaton with the multifaceted qualities of human behaviour, individual
reason, or sentiment.

Perhaps unsurprisingly, the famed automaton-showman Johann Maelzel became central to this C16.P35
nineteenth-century discourse, even in the highest political arenas (Register of Debates in Congress, XIV,
1
1837, col. 1441). In particular, his automaton trumpeter readily displayed the oppositions between
2
volitional human activity and automatic insensibility. As a new, wondrous, and somewhat inexplicable
clockwork thing, it initially elicited incredulous descriptions. Nevertheless, the trumpeter’s abilities were
p. 311 con ned to the expressions only pinned barrels, gears, and springs could attain. The automaton’s
premier London performances highlighted Maelzel’s wonderous mechanism alongside the machine’s
musical failings. According to an 1818 The Times review, the trumpeter lacked the nuanced musicality
customarily heard in a human performance. Typical of the hyperbolic eighteenth- and nineteenth-century
3
reportage on novel technologies (Decremps 1785: 69), the British reviewer proclaimed:

Nothing can exceed the [automaton’s] accuracy and neatness of the execution, or the steadiness of C16.P36
the tone: in the rapidity with which the same note may be repeated in succession, and in some
passages of a similar nature, it surpasses the powers of any living trumpeter; it fails only in
expression, and in the swell of the note, a defect which is common to all music produced by
mechanism.

4
4
(The Times, 12 Sept. 1818, p. 2)

The trumpeter’s repertoire, a selection of marches and fanfares, did not require any deep interpretive C16.P37
insights. It played strict military music with mechanical precision, steadiness, power, and speed, but not
rhythmical expression or intention. Unsurprisingly, the machine excelled exclusively in automatic qualities.

Maelzel’s Panharmonicon, another fascinating technology, was also a ruse of musical ability (Thayer 1921: C16.P38
5
251). Despite its grandeur, the machine was basically an ostentatious barrel organ, dubbed ‘a mammoth
music box’ by C. G. Hamilton (1916: 5). This clockwork-activated band of wind and percussion instruments,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
which Leonard Maelzel probably built in 1804, was proof in 1852 that ‘a full orchestra of clock-work
musicians is quite possible’ (‘Wonderful Toys’ 1852: 103). The Panharmonicon’s repertoire, much like the
trumpeter’s, focused almost exclusively on military music. Beethoven’s Wellingtons Sieg, the machine’s
best-known work, is arguably the composer’s most lacklustre. Critics should not fault the composer:
6
Beethoven set the work with the automaton’s limited musical abilities in mind.

In 1839 the Gazette Musicale noted the Panharmonicon’s ‘expressive play’, ‘extreme agility’, and ‘graceful C16.P39
manner’ (‘The Violin’ 1839: 270). But for those who had heard the rhythmical nuances and tempo
elasticities of a living, responsive orchestra, Maelzel’s barrel-organ contraption held no semblance of
musicality. Despite the wonder of it all, the technology contained little substance. ‘Let the rst novelty wear
o , and it will be seen by everyone, that the whole is a mere burlesque on the very name of musical
execution’, railed an American reporter in 1825, ‘Every man who wishes well to music, ought to be indignant
against such pretentions’ (Panharmonicon 1825: 136). Harmony, instrumentation, and timbre could be
mechanically replicated through Maelzel’s technology, but automatic parts could not replicate truly living
performances in time.

As prevailing nineteenth-century reactions to Maelzel’s machines reveal, clockwork musical behaviour was C16.P40
rarely something to lionize. ‘Compared with [an experienced pianist]’, exclaimed the New York Review
(1838), ‘how dead, how destitute of interest is mechanical music, even the wondrous melodium of Maelzel!’
p. 312 (‘Gardiner’s Music of Nature’ 1838: 46). Claims that the trumpeter produced tones ‘even fuller and richer
than those got out of a trumpet by human lungs and lips’ (‘Wonderful Toys’ 1852: 103) did not last long.
When Maelzel’s mechanical herald crossed the Atlantic, so too did its reputation as a musically dull
mechanism. George Templeton Strong’s critique of a human singer in 1842 is telling:

As to Mr. Charles Braham, his voice is good and he manages it well, but Maelzel’s automaton C16.P41
trumpeter has full as much expression. He looks as if he were some great piece of clockwork wound
up before the commencement of the concert, and made to work itself into the room and emit
musical sounds and then stalk out again at intervals.

(Lawrence 1988: 161)

Strong found the lacklustre performer to be ‘some great piece of clockwork’, a technology shared with C16.P42
Maelzel’s trumpeter, Panharmonicon, and his diminutive musical timekeeper. As Strong’s piercing
comment suggests, clockwork automata contravened the movements of the natural world, and living
motion itself. As this study will continue to show, Maelzel’s many clockwork devices seemed to project an
especially unnatural mode of performance activity (Hamilton 1898: 4), while promoting entirely insensible
7
and unmusical behaviours.
Interpreting Maelzelʼs Manufactured Musical Times C16.S3

Historical descriptions reveal the common lineage found in all self-moving machines: ‘Clocks, watches, and C16.P43
all machines of that [self-moving] kind, are automata’ (An Historical Miscellany, vol. 3: 281) and the
‘automaton’, was ‘a self-moving machine, without life’ (Thomson et al. 1836: 355). Clock-time, as
nineteenth-century society well knew, was an arti cial temporal construct set apart from the natural (and
empirically variable) daytime hours, the phases of the moon, the tides, and so on (Whitrow 1989). The sun

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
does not move to the case-clock. The tides do not move to the chronometer. By extension, the musician
naturally does not (and should not) play directly to the metronome-click. Richard Storrs Willis, writing in
The Musical World (1853), observed:

There are many persons, however, who mistakenly think that the intention of the metronome is to C16.P44
have its unvarying beat followed throughout an entire piece, denying all freedom to the play of
feeling. (p. 82)

Automatic performance activity, which included automatic musical movement, was largely detested in C16.P45
nineteenth-century performance practices. Indeed, it was a century in which even the simple pendulum—a
silent, non-automatic tempo reference—could be misused as an overly continuous beat-machine. For
example, William Ackermann, the inventor of a tempo-pendulum, inaccurately named the ‘Musical
p. 313 Regulator’, admitted in 1812, ‘it will be found to keep its motion much longer than necessary’ (Bingham
2012: 4). Nevertheless, Ackermann attempted a bizarre experiment for his day: he tried to play in tandem
with the pendulum movement. His report encapsulates the rhythmic aesthetics of a pre-metronomic
musical culture:

To play a whole piece strictly in time [with a silent swinging pendulum] has, however paradoxical C16.P46
the assertion may appear, upon late trial, purposely instituted for the sake of experiment, been
found to be attended with no good e ect. The music [performance], by the punctual [precise]
observance of time [as seen in a silent swinging pendulum], became divested of its spirit, and
highly insipid. (p. 4)

Those who understood the limited and nite applications of all tempo-machines often favored simple, C16.P47
silent metronomes (Encyclopedia Britannica 1884: 207). Swinging indicators such as Thomas Light’s
‘Harmonic Pendulum or Time-Gage’ (1825), were not intended to be sonic beat-dictators of performance
practices. The inventor’s description states as much: ‘The Pendulum when put in notion [sic] by a moderate
impulse, continues to [visually] vibrate equably, until it ceases to move altogether […] A few minutes [i.e.
moments] is a duration su cient for every required purpose’ (Bingham 2012: 94). Daniel Schole eld’s 1851
tape-measuring ‘portable metronome’ instructs its user to ‘hold the tape betwixt the nger and thumb, or
insert it between the pages of a book’ to view ‘the time intended’ (p. 181).

It was a di erent age of musical-metronomic precision to be sure (Landes 2000). Before national time C16.P48
zones and global standards for synchronization (Bartky 2000), nineteenth-century musicians quickly
recognized that clockwork metronomes poorly re ected prior, more vital modes of musical temporality.
Perhaps the earliest published critique of Maelzel’s tempo-apparatus appears in October 1817:

The constant loud ticking which it makes at every beat, though perhaps esteemed an advantage by C16.P49
some, who cannot measure equal portions of time in their mind, is disagreeable to those who have
a real feeling for music, and will render those who use it constantly, too mechanically uniform in
their performance, as it will not permit that judicious acceleration and retardation of the time
according to the genius of the passage, in which a great deal of the expression evinced by a
performer of taste consists.
(Philharmonicus 1817: 223–224)

Writing under a nom de plume, this musician upholds the early-modern principles of musical-temporal C16.P50
agency: a performer’s sense of time was to be prioritized above objective, mechanical measures.
Philharmonicus describes a common performance practice in which tempo is an expressive characteristic,
not an undesirable, mechanically driven process. As contemporaries con rmed (Czerny 1838: 69), Italian
tempo terminology such as adagio, andante, allegro, and presto were subjective sentiments of musical
motion before they ever transformed into metronomic onsets (bpm). Natural musical temporality was not

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
to be controlled by clockwork constructs. Czerny illustrates this non-metronomic modality of historical
p. 314 tempo with his own ‘Andante’ composition (Figure 16.2), for which ‘there are an in nity of cases, in
which a passage or piece may be played with several kinds of expression in respect to the degree of
movement’ (1839: 32).

Figure 16.2

Czernyʼs four di erent renditions of a four-bar ʻandanteʼ composition. In this brief passage, he reveals how musical temporality C16.F2
involves active rhetorical decisions, not passive attendance to mechanical-rhythmical regulations. NB: no interpretation is
labelled either ʻtempo rubatoʼ or ʻtempo giustoʼ, and no metronome mark is printed (Czerny 1839: 32).

Philharmonicus’ comment, which corresponds to Czerny’s directions, should not be mistaken as an C16.P51
extravagant call for rubato, or some new, romanticized performance a ectation. Neither musician describes
a Romantic ‘theft of time’. They advocate for a non-metronomic—or rather a pre-metronomic—
conception of nuanced musical movement. Their ideal musical tempi were rooted in anti-automatic
performance practices (Bonus 2013; 2014; Brown 1999).

Many others throughout the century observed this incongruity between auto-metronomic rhythm and a C16.P52
pre-existing musical temporality. Indeed, the so-called Romantic Age of musicianship did not usher in a
newfound love of clockwork tempo regulation. Summarizing the metronome’s many known negative
attributes, the 1884 Encyclopedia Britannica discounts its importance to musical pedagogy and practice.
According to this widely respected reference, Maelzel’s invention ‘reduced to mere mechanics what
formerly rested wholly on the performer’s feeling’ (Encyclopedia Britannica 1884: 207). For Gottfried Weber
(1779–1839) as well, the machine’s unnecessarily arti cial construction rendered it an incessant nuisance
8
rather than a trustworthy tool (Weber 1841: 75). More signi cantly, the clockwork metronome failed due to
a false premise—automatically produced MM beats could not project the realities of historical musical
motion:

The value of the machine is exaggerated, for no living performer could execute a piece in unvaried C16.P53
time throughout, and no student could practice under the tyranny of its beat; and conductors of
music, nay, composers themselves, will give the same piece slightly slower or quicker on di erent
occasions, according to the circumstances of performance.

(Encyclopedia Britannica 1884: 207)


p. 315 Maelzel’s technology is treated here as something of a passing technological fad. (As documented below, C16.P54
this nineteenth-century fad eventually became a twentieth-century temporal foundation). Indeed, for
many advanced artists, metronomic data proved to be super uous if not misleading information. ‘Such
indications [of tempo] can be found only in the feelings of the performer, or of the director’, realized Carl
Maria von Weber (1786–1826), ‘if they exist not in one of the two, the metronome is unable to supply the
want; all that this can do is, mechanically to prevent any gross mistakes’ (Weber 1827: 220). Characteristic
tempo terms were more useful, reliable, and appropriate markers of musical movement. As in uential
theorist and pedagogue Adolf Bernhard Marx (1795–1866) recognized, ‘The vague but less restrictive

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
indications by means of general terms appear to be more congenial than a rigorous subdivision into minutes
and seconds by the metronome’ (1853: 85). In 1847, a Harbinger critic expressed what seemed to be the stark
truth of the matter, ‘No strict time-keeping by Maelzel’s metronome can possibly produce a piece of music
as it existed in the composer’s mind’ (‘Musical review’ 1847: 185).

Rejection of the technology could be expected, because the metronome’s temporal modality was not an C16.P55
immediately understandable construct. Even today the bpm system is neither instinctual nor intuitive: all
amateur and professional musicians require training to associate their rhythmical actions and readings of
9
rhythmic notation with a constant machine-beat. (Examples of these modern metronomic pedagogies are
addressed later in this history.) Despite Maelzel’s early advertised claims, his metronome was not a quick
success. On 19 April 1818, he wrote in frustration to his business collaborator Beethoven:

There are stupid and lazy people who must be fed the truth with a cooking ladle, and who do not C16.P56
want to take any, not even the least trouble to learn something—and there are only too many of
these in Paris.

(Albrecht 1996: 137)

Maelzel’s wondrous and weird clockwork exhibitions were one business. The showman knew how to C16.P57
entrance audiences with occasional displays of unusual musical automata. But attempting to sell the
10
European public on an automatic appliance for music performance was another endeavour altogether.
Despite high-pro le published endorsements from some renowned musicians, who were initially invested
in the technology’s success—including Muzio Clementi, Samuel Wesley, Giovanni Battista Viotti, and
11 ,12
Frédéric Kalkbrenner in London (Attwood 1816: 3) and later Beethoven and Salieri in Vienna (1818) —
Maelzel could not fully convince contemporaries to employ automated tempo clocks throughout their
careers. Indeed, many of the original endorsers of the MM system listed here—including Beethoven—did
not follow up on their pledges to use it exclusively and comprehensively in place of traditional Italian time
words. In 1821, the Quarterly Music Magazine and Review noted with some chagrin, ‘We are sorry to observe
this promise is however ill-kept’, and that composers are ‘still too often content with marking their notes
p. 316 only with the common terms, and to leave the execution to chance or discretions’ (Quarterly Music
Magazine and Review 1821: 302).

Other contemporaneous composers considered Maelzel’s tempo-machine to be a useless conveyor of C16.P58


13
musical thought and motion. ‘As to an attempt to denote all the delicate shades of feeling, and the
consequent [metronomic] modi cations necessary to give full e ect to a performance’, Carl Maria von
Weber confessed, ‘I have found every endeavor fruitless, and have desisted from the task as hopeless’
14
(Weber 1827: 220). Ignaz Moscheles, an in uential ‘metronoming’ editor of Beethoven’s music, largely
concurred: ‘The player or conductor, who enters into the time and spirit of the piece must feel when and
where he has to introduce the necessary changes: and these are often of so delicate a nature, that the marks
of the metronome would become superabundant, not to say impossible’ (Schindler 1841: 111n.).

Beyond being insu cient markers for many nineteenth-century composers’ intentions, clockwork C16.P59
15
metronomes were not initially considered important tools for music instruction either. Maelzel’s earliest
advertisements describe the machine as a limited pedagogical proxy for a truly experienced and expressive
indicator of time (Parker 1825: 213). ‘The metronome also appears to us to hold out the greatest advantages
to young musical practitioners’, explained the 1816 British endorsement from composers and editors,
adding, ‘It serves as a complete guide to the pupil during the absence of a master’ (‘To the Editor’ 1816: 3).
The knowledgeable, sensible instructor remained the most accurate tempo-teacher throughout the
nineteenth century (Spohr 1833: 27). Indeed, the arti cial reference, according to some contemporary
artists and educators, created not ‘a steady timeist’, but something else: a thoughtless student who
mimicked rote, automatic behaviour. Under no circumstances could its clicks turn students into skilled

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
musical artists. In his Grand Violin School (1833), Louis Spohr’s instructions to teachers and students were
plain:

Exercises are to be repeated till the Scholar [i.e. student] is enabled to play them […] particularly in C16.P60
correct [musical] time. His success in the latter, the master may try by letting him here and there
play, to the beats of the metronome, but not too long, as otherwise the playing soon becomes sti
and awkward.

(Spohr 1833: 39)

Traditional nineteenth-century music educators seldom applied automatic sound-beats as a guiding, rst C16.P61
principle of musical temporality. Spohr imparted the standard pre-metronomic lesson: beat-technology
was an optional, supplementary reference. Although it could serve some limited, meagre assistance, the
machine posed signi cant dangers when used for ‘too long’. In 1836, The Musical Magazine reiterated that
teachers must take care with the device: ‘We would not recommend its constant use […] but occasionally
introduced it will be of great service’ (‘The Metronome’ 1836: 304). If the automatic metronome was to be
referenced at all, discretionary treatment was considered a necessary condition of its pedagogical
employment. In 1886, the systematic American educator A. R. Parsons attempted to quell fears about his
p. 317 own regulative musical-gymnastics pedagogy with the rhetorical question: ‘Does any one still hold the
use of the metronome to be dangerous to musical sensibility?’ (Parsons 1886: 24).

As Parsons’ query suggests, many during his age continued to answer a rmatively: yes, the metronome C16.P62
indeed hampered long-term musical growth. The critical discourse relates the radical disservice done to
musicians who faithfully used Maelzel’s clockwork metronomes. For so many pre-modernist musicians, the
technology did not reference a composer’s desired tempo. Rather, it mechanized musical time beyond any
and all creative intentions, because it constricted the performer’s rhythmical agency. ‘It must be borne in
mind that Beethoven’s instrumental music has undergone a metamorphosis’, Anton Schindler re ected
after the composer’s death; ‘it is necessary rst to acquaint the reader that this metamorphosis relates
16
wholly and solely to metronomising, or the regulation of time by means of the metronome’ (1841: 97).

As an arti cial device opposed to performer artistry, the metronome accrued a pejorative reputation—one C16.P63
identical to Maelzel’s other self-moving machines. Music students and professional performers were heard
transforming into mindless marionettes when moving synchronously to the invariable pulse of the
17
clockwork metronome. If mistakenly employed as a rhythm regulator, ‘for the mere purpose of learning to
keep time’, the metronome ‘would take away the spirit of the pupil’s performance, making it a mechanical
a air, and him, if he succeeds in conforming to it, a slave of time’—this according to The Musical Magazine
in 1840 (‘Substitute for Maelzel’s Metronome’ 1840: 67). The archetypal automaton as unthinking puppet,
18
lacking in vitality or volition, is recounted in many similar assessments (Cooke 1917: 213). The metronome
emblematized, as with Maelzel’s clockwork trumpeter and Panharmonicon, shallow and dumb ways of
19
being (Elson 1897: 299–300). ‘To bang through an overture like a machine is not the thing’, The Harbinger
chided in 1846:
No machine has whims and in ections, and therefore, it only makes cast-iron music […] a mere C16.P64
beater of time is worth nothing, but to embarrass all parties and to kill the music, and Maelzel’s
metronome were quite as good a thing and less expensive.

(ʻMusical reviewʼ 1846: 204)

The clockwork metronome’s regulative sound, as described throughout the century, developed musical C16.P65
workers or slaves, not rhythmically sensitive musicians. Failing to project the subjective, sensory nature of
musical temporality, Maelzel’s metronome—identical to its automatic brethren—more accurately

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
projected the stark mechanical time and motion that propelled the Industrial Age (‘Listening to Music’
1897: RB2).

As suggested by evidence presented later in this chapter, neither the jab ‘like a metronome’ nor the C16.P66
pejorative ‘metronomic’ nd much space in modernist practices which focus upon recreating the most
mechanically accurate tempi, objectively precise rhythms, or authoritative bpm solutions to Beethoven’s
compositions. Yet the nineteenth century was awash with metronomic insults (‘The Symphony Society’
20
1894: 2). This wealth of historical invective—be it colourful metaphor or realistic observation (‘Music’
p. 318 1856: 115)—attests to the historical view that playing like an automatic machine, and playing with the
21
metronome’s click, was once musically abhorrent (‘Herr Stavenhagen’s Recital’ 1891: 3).

Yet, given the showman’s reputation as mechanizer of musical performances, it seems logical that Maelzel C16.P67
was a orded a parallel—and positive—cultural reception in industrial venues, which grew to value e cient
and systematized labour practices. Indeed, while many artistic musicians dismissed his metronome,
Maelzel was conversely lauded for inspiring a newly automating age of factory production. In 1836 Senator
Daniel Webster (1782–1852) described Maelzel’s automata as archetypes for America’s rapidly
industrializing landscape:

These automata in the factories and the workshops are as much our fellow laborers, as if they were C16.P68
automata wrought by some Maelzel into the form of men, and made capable of walking, moving,
and working, of felling the forest or cultivating the elds.

(Webster 1903: 69)

The politician envisioned a nation where arti cial automata and people worked with the same movements C16.P69
and purposes; where human labourers were functionally, if not visually, identical to Maelzel’s machines.
Webster was not alone in nding automata’s superior trait—that of precisely repeating a series of limited
actions—to be a prime exemplar for worker behaviour. In The Philosophy of Manufactures (1835),
industrialist Andrew Ure envisioned a modern labour system, in which humans performed as an automated
ensemble, one no less impressive than Maelzel’s Panharmonicon (Ure 1835: 13–14).

Ure knew that the Maelzelian technologies were available to create mass production systems, but people’s C16.P70
working methods were not yet up to these automatic standards standards. His ‘main di culty, did not […]
lie so much in the invention of a proper self-acting mechanism [but], and above all, in training human
beings to renounce their desultory habits of work, and to identify themselves with the unvarying regularity
of the complex automaton’ (p. 15). For Ure, the greatest problem in establishing the new industrial factory
was in making individuals habitually metronomic in their thoughts and actions.

It might seem self-evident that labour theories and artistic performance practices were two opposing value C16.P71
systems in the nineteenth century. Ignacy Jan Paderewski (1860–1941) reiterated the long-standing
distinctions between clockwork action and innate human behaviour: ‘To be emotional in musical
interpretation, yet obedient to the initial tempo and true to the metronome, means about as much as being
sentimental in engineering. Mechanical execution and emotion are incompatible’ (Finck 1909: 455).
For those who heard living music made automatic, Maelzel’s metronomes did not enhance skill; they did C16.P72
not hone artistic abilities; they did not de ne the right tempi—by Beethoven or anyone else. ‘We cannot
depend on metronome tempi’, concluded the piano virtuoso Fannie Bloom eld Zeisler (1863–1927), ‘for
they are not reliable’ (Brower 1917: 185). Yet, as Paderewski’s critiques suggest, a new obedience to
metronomic data and its automatic rhythm was ascendant. The pianist-composer believed the ‘tempo
p. 319 sticklers and metronome believers’—the new technologically progressive pedagogues—were focused on
the wrong temporal epistemology; motivated by an arti cial modality of musical action. As he and other
traditionalists argued, the metronome-believers, in seeking mechanically objective answers to pre-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
mechanistic musical expressions, ultimately preoccupied themselves with misdirected, industrially inclined
performance outcomes (Sternberg 1917: 82).

Nevertheless, the Industrial Age, once only dreamt of by Ure, was on the march; and with it came the C16.P73
ascendance of both mechanized working methods in factories and metronomic pedagogies for musical
performance. As this history continues to chart, the ideals and practices of musical tempo were changing,
just as the values underlying social time-measurement were changing. More precisely, the communal times
were becoming ever more automatic, and ever more arti cial, in the city, the factory, and the music room.

Modern Metronome-Believing: Automating the Musical Times C16.S4

It was during Paderewski’s lifetime when Maelzel’s automatic measurement system began to rede ne C16.P74
22
traditional models of musicianship and reorient past pedagogical aims (Bonus 2017). Concurrently,
Maelzel’s once magical automata became the stu of history. They were e ectively replaced by utilitarian
factory machines, a ordable household clocks, and personal watches, all of which now helped to drive
modern society ever forwards. During this pivotal historical moment, the metronome was reconsidered as
an essential apparatus for the development of good, accurate musical behaviour. For a culture swayed by
Industrial Age ideals of time and motion, acting more automatically and precisely in time was becoming an
23
imperative, not only in labour practices but in modern musicianship as well.

From 1890 until 1920, this ‘metronomic turn’ reveals itself through pedagogical and critical discourses in C16.P75
con ict. For those cognizant, like Paderewski, of pre-metronomic traditions, the values of musical artistry
were undergoing a radical cultural upheaval. They keenly observed the metronomic turn in action, as
technologically dependent pedagogies and practices ascended, and as prior temporal qualities—including
willed elasticity, metric accentuation, and rhythmopoetic in ection—were being dismissed. ‘Metrical
regularity has seemed so important to many students of interpretation’, lamented Henry G. Hanchette in Art
of the Musician (1905), ‘that they insist upon steadiness of time-keeping and exactness of beat-recurrence
even at the expense of much that might otherwise be added to expression in performance. Hence we nd the
schools and the pedagogues demanding metronomic accuracy and mechanical precision of pulse’ (p. 29).

In 1908, Vassar College professor Kate S. Chittenden (1907) similarly observed metronomic-musical C16.P76
behaviour becoming central to the latest educational methods: ‘Piano-teaching has changed so much in the
p. 320 last fteen years that grade work in one sense is not so hard to arrange as formerly; that is, if one be
24
governed by technic exercises played with the click of the metronome’ (1907: 144). Metronomic sound
simpli ed modern pedagogy by regulating tempo and reducing movement to reproducible certainties. The
automatic system made teaching musical time to the masses easy. But this ease and e ciency came at a
noticeable cost. These modern pedagogies—increasingly distanced from the culture and aesthetics of
25
Beethoven, Brahms (C.A.B. 1888: 850), Paderewski, and Rachmanino —were, according to Chittenden,
‘built rather too much on the dangerous metronomic basis of mere exercises which develop machines’
(1907: 146).
Such warnings went unheeded by progressive pedagogues. Precise clockwork timekeeping, as an C16.P77
increasingly normative temporal construct for modern society, logically transferred to precise tempo-
keeping. Metronomic memorization procedures, such as those found in James Matthew McLaughlin’s
Elements and Notation of Music (1902), imprinted on young performers the rhythmical fundamentals that
26
were intertwined with clockwork technologies (1902: 80). In Half-Hour Lessons in Music (1907) the
impressionable novice learns:

Now time in music, children, is expressed by steady, even counts, or beats […] We reckon time by C16.P78

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
the clock in seconds and minutes, and the metronome ticks in exactly the same way as the clock
[…] Each tick of the clock means one second—each beat of the metronome means one quarter.

27
(Kotzschmar 1907: 13–14)

Such mechanistic musical practices migrated to even more advanced pedagogies in the following decades. C16.P79
Elson’s ‘Modern Piano Methods’ (1918) provides archetypal Machine Age instructions, whereby: ‘The slow
trill of two notes is repeated thirty to forty times with each pair of ngers, at a metronome rate of 40 to 60
28
for each note’ (p. 70), while ‘broken chords […] should be repeated from ten to twenty times, with the
metronome’ (p. 72–73). Not merely a convenient analogy for children’s lessons, the clock-time/music-time
convergence appears in sources as divergent as a 1922 treatise for pianola manipulators (Grew 1922: 32) and
Gustave Langenus’ popular handbook Rhythm-Builder (1933):

The use of a Metronome will greatly help in acquiring a good rhythm […] Rhythm after all is more C16.P80
or less [an] arithmetical problem, which once solved, enables one to concentrate more easily on
reading notes, ngering and phrasing […] the clock is again a good illustration […] Musically
speaking, the ‘tick’ is the accented beat, the ‘tock’ the unaccented. (p. vi)

In this modern era of musicality, clockwork metronomes were essential apparatuses for building time and C16.P81
‘acquiring a good rhythm’. The American educator E. M. Bowman attested, ‘The metronome will give you an
exact standard of metre and rhythm, and help to train your sense of time’, with the justi cation: ‘The
p. 321 metronome is a cold-blooded machine. It works the same way every day. It never gets excited or
discouraged. It does not balk or run away’ (1910: 155).

Cold-blooded automata eventually found an era in which their measurements and movements were deemed C16.P82
virtuous models for musical functioning. ‘I cannot urge too strongly the constant e ort to play in time’, said
Harriette Moore Brower in Self-Help in Piano Study (1920), upholding the new techno-dependent pedagogy:
‘We must have a just sense of the mathematical values of notes. This is only acquired by constant
timekeeping and counting’ (Brower 1920: 16–17). It was a Maelzelian magic trick that took over a century to
29
uncover: musical rhythm became analogous to automatic performance activity (Franz 1964: 29–31). By
the Machine Age of the 1920s, the metronome pulse had clearly morphed from the abhorrently arti cial to
30
the absolutely accurate (p. 38). No longer a subjective guide, as Moscheles once assumed, and no longer an
optional, limited indicator for more vital movements, automatic sound-beats became central to a new mode
31
of musicality (Brower 1920: 18).

This mechanical-temporal metamorphosis—so readily recognized and derided in the previous century— C16.P83
32
continued to be intuited by some, yet its root causes were often being obscured (Martin 1929: 108). For
traditionalists, modern metre was conceptually and practically being severed from vital origins (Salmon
33
1920: 587). The sublimation of a fundamentally innate, historical musical temporality might be predicted
if modernist lessons, such as this one from 1925, were accepted on faith:

PAY ATTENTION TO THE BEATS C16.P84


However, the student must be careful to understand that keeping steady time means that each note C16.P85
of the exercise must coincide with each note of the metronome, and not simply to play on and on
while the metronome keeps ticking, each at variance with the other. Such practice is valueless.

34
(Franz 1964: 39)

Throughout the century, modern pedagogies of all sorts continued along these metronomic lines. The C16.P86
prescription was unambiguous: it is a social virtue to be more automatic, more systematic, and
35
correspondingly, less individual in rhythmical thought and action (Hall 1907: 4–5).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
The distinction is worth reiterating: these were not the temporal ideals of nineteenth-century musical C16.P87
artists. As I have charted elsewhere, the striving for metronomic-rhythmic behaviour—once artistically
repellent—was rst championed by late nineteenth-century experimental psychologists (Bonus 2014;
2017). Their metronomic performance practices were disseminated throughout modern society via a host of
popular publications, pedagogical prescriptions, and especially the testing procedures known as child study.
As child study illustrates, the sociocultural values of time and action were increasingly being de ned by
automatic standards. At the beginning of the twentieth century, metronomic performances were evident
throughout modern society: in schools, factories, laboratories, and other arenas where musical expression
p. 322 was of little concern. The modern act of accurately performing to an automatic tempo (i.e. ‘mechanical
execution’) began in earnest as a scienti c-industrial procedure, not an artistic imperative (Figure 16.3).

Figure 16.3

Detail from ʻTests and Measurements in Child Studyʼ. To gauge a childʼs physical endurance, the experimental psychologist C16.F3
connected the subjectʼs arm to a standard chronographic apparatus, the ergograph. The procedure required this subject to
continually li his bound index finger to the adjacent metronomeʼs constant beat (Marshall 1901: 420).
The Current Mechanics of Musical Time C16.S5

As nineteenth-century reception history suggests, none of the constituent elements of historical musical C16.P88
temporality derived from the metronome beat. Nineteenth-century composers’ tempo modality did not
originate in modernist precision-time constructs, or scienti cally imbued training methods, or
metronome-dependent rhythmic theories. Past musical principles seldom if ever aligned with industrial-
era penchants for automation. Beethoven’s supposed nal words on the matter—‘No metronome at all!

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Those who have a right feeling do not need it, and those who have not, will not be helped by it’—echoed
p. 323
across the century. Regardless of whether or not the statement was original to Beethoven, Liszt believed it
to be authentic, as did a host of skilled composer-performers (Kullak 1901: 22; Liszt 1894: 75). Marx attested
to ‘the unanimous declaration of Czerny, Madame von Ertmann, and others [that] Beethoven […] played his
compositions di erently every time’—a recognition that made metronoming the composer’s publications
either futile or highly misleading (Marx 1895: 68).

It is also worth recollecting Hector Berlioz’s tempo aesthetics, which encapsulated the prevailing temporal C16.P89
attitude of nineteenth-century performance culture. Berlioz cautioned, ‘I do not mean to say […] that it is
necessary to imitate the mathematical regularity of the metronome; all music so performed would become
of freezing sti ness’ (Berlioz 1856: 246). The pre-modernist age was never so invested in the precise
clockwork regulation of vital musical activities. Beethoven, Berlioz, and Brahms were not labour theorists,
time-study scientists, or experimental psychologists; and their embodied approaches to musical pulse,
gesture, and metre cannot aesthetically be pinned to an automatic framework of time and motion (Figure
16.4).

Figure 16.4

Depicting e icient, metronome-defined ʻrhythmicalʼ behaviour in twentieth-century labour practices. From the pro-industrial C16.F4
film Conquer by the Clock (RKO-Pathé, 1943); still image at 1 minute, 51 seconds. Shown here is the prescribed metronomic ideal
for productive human behaviour. The tableau ideologically corresponds to contemporaneous time study experiments, musical
pedagogies, and metronome-reliant compositional practices.

p. 324 It cannot be denied that automatic metronomes are valuable and, in many cases, essential to many modern C16.P90
musical practices, which stipulate the enacting of decidedly non-sentimental types of rhythmic action.
Since the twentieth century, certain progressively minded composers have embraced an automatic-
temporal aesthetic, one which demands high degrees of metronomic precision in practice and
36
performance. But in uential pedagogues, performers, and tempo scholars should not presume that skilled
musicians from the past valued metronome technologies with the same faith and fervour that many current
metronome-users do.

The meanings of musical motion, and the ways musicians measure those motions, have changed over the C16.P91
centuries. And it must be acknowledged that modern, ‘classical’ musical culture has in a great many
instances altered in sympathy with the latest metronomic quanti cation systems. Historical music

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
performance can indeed be swayed by scienti c-industrial aims and apparatuses. In Rhythm and Tempo
(1953), the musicologist Curt Sachs measured his heartbeat to the metronome’s tick, presenting his
ndings as a basis for historical tempo (Sachs 1953: 33). Further guided by this bpm mentality, Sachs made
the anachronistic claim, ‘ours is a mathematically counting notation’ in which ‘the quarter note is our
motor unit’ of musical time (pp. 168, 173, 201). Sachs’s modernist conceptions of musical temporality,
similarly shared by contemporaries Jaques-Dalcroze (1865–1950) (Bonus 2017), Stravinsky, Joseph
37
Schillinger (1895–1943) (Encyclopedia of Rhythms 1976), and Scherchen (1950: 87–88), to name only four,
are revelatory for reasons that go beyond their exacting rhythmic and metric calculations (Bonus 2014).
Each music professional points to a profound recalibration in the meanings and enactments of musical
38
time, which cannot be traced to Maelzel’s original intentions (Rothschild 1961: 102) or Beethoven’s own
39
compositions (Schindler 1841: 104–105, 116–117) (Figure 16.5).

To perform with precise ‘motor units’ of time was once an act reserved for science and industry. Over a C16.P92
century ago, metronome-wielding psychologists—desiring the most e cient and accurate working
(‘business’) practices—established a need for the mechanical time study of musical endeavours (Bonus
2017). In 1915, Carl Emil Seashore—directly in uenced by the previous generation of experimental
psychologists—suggested conservatory musicians undergo chonographic study, to gauge their ‘precision in
the time of rhythmic action’ (Seashore 1915: 146; 1919: 203–204). These scienti c values would eventually
reach the musical mainstream. Bartók’s publications often included both stopwatch time studies and highly
speci c metronome indications, to precisely set out his desired performing tempos and durations. It is
worth noting that Bartók—for whom enforcing metronomic rhythm in pedagogy, composition, and
performance became a musical imperative—was criticized in 1922 for playing ‘with the sti ened metal
muscle of a jerkily rhythmic automaton’ (Philip 2004: 174). Yet, in 2008, the mechanical time study (and by
extension time training) of musicians seemed an unradical proposition, when New York Times reviewer
p. 325 Bernhard Holland suggested ‘conservatories […] hire time-and-motion experts, professionals who could
point out that the ailing arm, the bulging eye and the balletic upper torso are extraneous work in a
business best devoted to doing the most with the least’ (Holland 2008). Precise clockwork tempi and
mechanical time controls are no longer so distant from human creative endeavour or agency. No longer an
aberration, the metronomic musician can be considered a regular, if not a positive, exemplar of our modern
40
performance culture (Osborne 1989: 96).
Figure 16.5

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Tracing the origins of the automatic modality of musical rhythm—and the automation of musicianship—through the C16.F5
experimental psychologistʼs metronomic-chronographic methods and aesthetics. The hash signs represent the normal
metronomic onset, the so-called ʻregular accurate rhythmʼ. The dots mark the subject-performerʼs ability to conform, during the
laboratory experiment, to the automatic beat standard (Scripture 1895). The theoretical and practical oppositions to Czernyʼs
tempo lessons are readily apparent in Scriptureʼs metronomically measured values of ʻrhythmic actionʼ.

p. 326 Paderewski’s tempo sticklers and metronome believers are still with us, and they have even more precise, C16.P93
41
automatic, and physically invasive measuring tools at their disposal (Miller 2007: 214–215). It must be
recalled that A. B. Marx found G. Weber’s home-made tempo pendulum ‘[a] more simple contrivance, which
is cheaper and less liable to get out of order than the wheelwork of Maelzel’ (1853: 83). Indeed, other
nineteenth-century endorsers of simple tempo-pendulums strongly suggested that neither clockwork
precision nor automatic sound impressions were vitally important to the notions of musical temporality. Yet
the metronomic click-track currently emitted by phones, computer programmes, tuners, and so on—being
entirely detached from visible space and natural motion—has further untethered musical rhythm from
historically expressed, vital origins. As described by current advertisements and patents, complex tempo
machines have become workaday hegemons over all musical interpretations—and consequently over all
performers’ rhythmical movements. A hermeneutical reading from a recent metronome manual may su ce
to uncover the automatic ideals motivating so much of modern musical timekeeping:

Practicing to a metronome is foundational to understanding [believing in automatic] rhythm and C16.P94


developing [entraining] one’s inner musician [physically and psychologically] […] The Soundbrenner
Pulse is a tactile [physically automatic] alternative to audible [sonically automatic] metronomes and
click tracks, enabling live [living] performers and session players to stay keyed in to their
environment without losing [deviating from the automatic] time [technology].

(Sweetwater 2018)

Although this particular machine seems somewhat novel (U.S. Patent No. 1,783,537, Goudsmit 1930), the C16.P95
promotional claim is not. It follows the technocultural currents of twentieth-century musicianship, which
often charge performers to play synchronously with automatic metronomes in order to attain the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
presumably accurate musical time. ‘Metronomes are well known in the musical art’, states Clair Omar
Musser in her 1965 U.S. patent ling:

Metronomes are arranged to provide a series of audible signals at a predetermined rate so as to aid C16.P96
a musician in keeping his musical rendition operating at a constant rate.

(U.S. Patent No. 3,263,551, Musser 1966)

Contradicting Philharmonicus’ claims against the very rst (and comparatively imprecise) metronomes, C16.P97
modern musical discourses often repeat the same lesson: the fundamentals of ‘timing and rhythm’ are no
longer centred in the self; they do not originate in the performer’s mind, will, expression, or pulse-sense
42
(Miller 2007: 214–215). Unlike novel metronomic lessons from the early twentieth century, the wall-clock
no longer provides a convenient analogue—the beat of the modern metronome speaks for itself. ‘Some
students complain that they have a hard time playing while the metronome is ticking’, a contemporary
instructor, oblivious to all historical theories and aesthetics, maintains: ‘that’s precisely why they need to
use it’ (p. 214).

p. 327 The precision-striving relationships between musicians and their metronomes continue. As research on C16.P98
this topic will continue to show, many modern meanings of musical time correspondingly promote
performers’ physical and mental acquiescence to automatic bpm references. Indeed, in the modern musical
era, the terms ‘beat’, ‘tempo’, and ‘rhythm’ continue to be used as metonyms for automatic behaviour
(Lamb 2014: 80–81). For countless performers today, to be ‘in time’ and to ‘have rhythm’ is to be precisely
metronomic. What is now being promoted as accurate musical tempo was once observed as dumb
automatism (Figure 16.6).
Figure 16.6

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
A typical college-level rhythm manual, which subversively illustrates the modern association between the musical beat and C16.F6
metronomic rates for ʻall music played, sung or heardʼ (Snyder 2001: 2). The heartbeat images mirror exacting automatic sound-
impressions. (Past musical notions of arsis–thesis pulsation have no place here.) To be more faithful to its underlying techno-
rhythmic ideology, the instructional text should replace the mechanically replicated heart icons with metronome images. This
modern musical theory aligns with scientific-industrial ideals of ʻaccurate rhythmic actionʼ, which were endorsed by
experimental psychologists starting in the late nineteenth century.

The quarter note was not an unstoppable ‘motor unit’ for all musical times. Yet many still consider the C16.P99
modern metronome beat to be the historical foundation of all musical movement. The misplaced
metronoming of musical time and action is not a problem relegated to past performers, or one described in
old and obscure sources. The technocultural anachronism can be witnessed today. Take for example a
February 2013 episode of Radiolab, a radio series and podcast devoted to contemporary issues of technology,
science, and culture (Radiolab 2013). The show exposes again how the metronomic beat-modality is
presumed to be the normative underpinning of pre-modernist musical time, even for some professional
performers and scholars.

The Radiolab segment attempts to justify playing Beethoven’s symphonies at ex post facto printed C16.P100
metronome numbers—using an anachronistic auto-sonic metronome as a strict tempo regulator. To
bolster their pro-bpm argument, the participants—the main host, a professional conductor, and a
p. 328 professional scholar-critic—quickly discount many past criticisms: that Beethoven had a broken
metronome; that he was deaf; that the acoustic practicalities of music performance disallowed one absolute
tempo rate.

Radiolab listeners are then o ered an uncritical, modern-day approach to metronome-believing, as they C16.P101
learn that Beethoven’s metronomic intentions, authoritatively rooted in his mind, deserve obedient
realization in the physical world. The conductor presents the epitome of the intentional fallacy: ‘If we can
create the music that Beethoven heard in his head, isn’t that something worth doing?’ Their rst mistake
was in presuming that a precise and continuous metronome click was integral to the music heard in the
43
composer’s head. The modern conductor’s question is taken further, out of fallacious history and into the
realm of practical application. The host summarizes his musical guests’ professional viewpoints:

Both [the professional conductor] and [the published scholar-critic] think that we probably should C16.P102
just accept these accelerated tempos, you know like the 5th [Symphony] at 108 [sonic bpms] […]
44
Just go with it.
The prescription could be restated: the modern metronome is the ruler of all musical time, a fact which C16.P103
justi es the sublimation of musicianship to its click; performer agency and historically expressed ‘artistry’
factor little into the qualities of nineteenth-century musical tempi. A revised aesthetic—one better suited to
the experimental psychologist’s laboratory—then justi es this modern process of musician mechanization.
As the scholar-critic explains, when musicians passively observe the published MM number (using their
modern metronomes), ‘The piece [Symphony No. 5 in C minor Op. 67 (1804–8)] is always feeling like it’s
running away from us […] In a very real psychological way’—just as Beethoven must have scienti cally
intended.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Not one of the collaborators broached the self-re ective question: was their entire enterprise—completely C16.P104
linked to an anachronistic digital metronome-signal—wanting in a very real psychological, aesthetical, and
practical way? No matter. With Beethoven’s psycho-metronomic motives rmly established, the Radiolab
participants then re-enacted, for the entire world to hear, not Beethoven’s expressive intentions, but a late
nineteenth-century work experiment—a century-old laboratory procedure that tested human tolerances
for endurance, attention span, and mechanically objective rhythmic action through increasingly rapid
metronomic rates. Their modern musical journey proceeds apace:

host: You’ve never heard [Beethoven’s Symphony No. 5] faster than 108 [sonic bpm]? C16.P105
conductor: I’d think you could do 120-ish C16.P106
host: Well let’s just do this. Let’s get out our [miniature digital] metronomes out. C16.P107
conductor: Like here’s 120…Click-click-click… C16.P108
p. 329 conductor: [Sings some bars of the exposition to the clicks] C16.P110
host: Ok. Well, make it faster. Make it 140. C16.P111
conductor: 140, I bet…Click-click-click… C16.P112
conductor: [Sings the same exposition bars to the clicks]…you can do it. C16.P114
host: [Sings the same bars]. Can you go to, like, 160? C16.P115
conductor: I think that’s around the edge… C16.P116
host: [To listeners]. But we tried it with his quartet [as a proxy for an entire orchestra.] C16.P117
With the digital bpm click-rate sounding, musicians play the exposition bars in synchronicity. They stop. C16.P118
Group: Yeah! Wow! [Laughs] C16.P119
Host: That was fantastic! You totally nailed 160! C16.P120
Conductor: [volume reduced] I don’t know if I’d say nailed… C16.P121
45
Host: [To listeners]. That is a Beethoven I could dig, right there! C16.P122

One could argue that this passage is o ered for the sake of rhetorical convenience, and that—beyond a C16.P123
popular, globally broadcast radio show—present-day Beethoven performances o er far more complexities
in regard to interpretation. Careful listening to the state of performance practices across the world will help
to determine the larger reality. Nevertheless, this event was an actual, automated musical performance,
made real through the minds and bodies of living performers. It documents a now familiar interpretive
approach, whereby an automatic sound-pulse regulates the motion of a pre-modernist musical
composition. The recorded act itself represents a serious reduction of Beethoven’s more complex
rhythmical intentions. Is this not the continued ‘metamorphosis’ through ‘metronomizing’ that Schindler so
derided? Is this not the ‘freezing sti ness’ of metronomic regulation that Berlioz warned against? But here,
in the current age of musical accuracy, to be ‘metronomic’ did not provoke any serious criticisms; no dire,
universally acknowledged claims of lifelessness; or observations of unthinking automatism. If anything, the
process evoked adulation. The very existence of this musical event exempli es the temporal aesthetics
made operative since the metronomic turn. Being musically in time—for so many lessons, practice sessions,
and rehearsals that go unrecorded—means to play with the metronomic ‘motor unit’, guided by an
Industrial Era faith in the virtues of auto-mechanical precision (Figure 16.7).
Although this brief survey cannot hope to document every instance of the modern, metronomic C16.P124
performance practice, evidence presented here is entirely indicative of its presence as an everyday, seldom-
46
questioned aspect of current musicality, and by extension musical temporality. The modern obsession
with Beethoven’s meagre metronomic indications, the search for his best, most correct MM number,
continues to distract many from the more relevant recognition: automatic metronomes support scienti c-
industrial desires—not to de ne the right, ‘sentimental’ tempi, metrical accentuations, and rhythmical
p. 330 gestures of the past, but to systematically regulate psycho-physical behaviours in the here and now.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Figure 16.7

Originally fig. 48 from the chapter ʻThe Art of Labourʼ (Amar 1918). ʻA Martinique soldier working with the self-registering C16.F7
Jointing-plane.ʼ An essential apparatus for psycho-physical testing and training, the clockwork metronome appears to the le .
Corresponding to progressive, twentieth-century musical practices, the automatic metronome is used as a laboratory apparatus
to gauge and regulate the entirety of a labour performance.

Perhaps today’s metronome believers should consider whether their zeal for metronomic precision creates C16.P125
the real, unsolvable conundrum in nding the right historical-musical movements. ‘It is the mark of an
educated man to look for precision in each class of things just so far as the nature of the subject admits’,
Aristotle stated in the Nichomachean Ethics, presaging the fallacy guiding scienti c-industrial tempo study;
‘it is evidently equally foolish to accept probable reasoning from a mathematician and to demand from a
rhetorician scienti c proofs’ (Aristotle 1959: I3, B12–27). Paderewski echoed this ancient truism with his
conclusion, ‘Mechanical execution and emotion are incompatible’. If these past observations—and all other
nineteenth-century criticisms presented here—are to be taken seriously, then perhaps today’s tempo
chronographers will forever misconstrue the sentimental tempi of the past when they rely upon the highly
precise clockwork controls of the present. According to pre-modern musical artists, metronome-believing
itself is the instigating fallacy. For those who boldly rejected the premise of Maelzel’s clockwork
metronome, the real problem in nding the right tempi originates from a misplaced desire to automate the
very fundamentals of human-derived musical motion.
p. 331
A Lasting Maelzelian Showcase C16.S6

Johann Maelzel’s career seemed dedicated to one vision: the replication of natural actions in the repetitive C16.P126
motion and sound of clockwork. For nearly a century, the reception and recollection of Maelzel’s machines
show them to be mere mechanical copies of more complicated, living creations. Observers continually
acknowledged that Maelzel’s trumpeter and Panharmonicon were limited clockwork facsimiles of musical
performers; while the cold, unfeeling chess player was an eerily human thinker (since, underneath the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
clockwork façade, he was human).

In Maelzel’s automata, many witnessed the possibilities of a fully mechanized future, especially in industry, C16.P127
where labourers could indeed act automatically through methodical and redundant training. Yet for
nineteenth-century artistic endeavours, automatic action was nothing to be emulated. A few perceptive
musicians recognized that Maelzel’s culture directly extended to the metronome, that automata and
47
tempo-clocks shared an identical father (Sternberg 1917: 79–80). The metronome, like other devices from
48
Maelzel’s exhibitions, reduced nuanced musical actions into shallow, clockwork simulacra. As so many
over the nineteenth century claimed, the metronome was nothing more than an automaton that engendered
automatical behaviour.

Despite these past realizations, self-moving technologies, with the values of systematization and C16.P128
reproducibility entailed to them, are seeded throughout modern life and musical activity in ways that might
have shocked the machine-showman. The acts of automatic beat dictation, tempo regulation, and
clockwork quanti cation—which had countered the musical sensibilities of Maelzel’s Parisian audiences,
c.1818—would eventually become social norms beginning in the twentieth century.

Implacable clockwork qualities once bemoaned as unmusical, dead, and arti cial have now come to C16.P129
motivate a great many musical endeavours. Indeed, it can be argued that Maelzel’s machine culture
prospers throughout Western society more now than ever. Automatic playing technologies are xtures in
the commercial and consumerist-music landscape. Step sequencers, drum and looping machines, along
with CD and MP3 players, can be recognized as the progeny of his once magical and bizarre clockwork
automata. Metronomic time, emanating from his once-detestable rhythm clocks, currently resounds with
even more precision and constancy through state-of-the-art electronic and digital technologies. And the
modern performer, more often than not, has learned to accept on faith Maelzel’s mode of musical time and
action. ‘When using the metronome to improve your sense of timing and rhythm’, a contemporary teacher
uncritically prescribes, ‘listen to the metronome tick for a while. Move your body to the rhythm of the ticks.
Once you’ve internalized the [automatic] rhythm, you’re ready to start playing’ (Miller 2007: 214–215).

Although an obelisk-shaped ticker evidently remains as the one lasting emblem of his ingenuity, the far- C16.P130
famed Johann Maelzel—that entertainer, entrepreneur, engineer, and trickster—bestowed upon the
p. 332 modern age much more than tempo machines. It took over a century, but he posthumously succeeded in
selling consumers on his two most radical technological notions: the rst, that human senses might be
attributed to mere, mindless automata; and the second, that arti cial clockwork constructs could be
imposed upon sensible humans through mere, automatic metronomes.

Notes
1. E.g. Virginia representative John Robertson, on 11 Oct. 1837, derided feckless legislators in the House of Representatives C16.N1
for being ʻslaves, mere puppetsʼ, indistinguishable from Maelzelʼs machines: ʻThose who desert their own principles, and
act in opposition to their own judgments, are slaves, mere puppets, moved by the will of another. Maelzel could construct
a House of Representatives as fit to exercise the functions of legislators—yes, sir, speaker, orators, and allʼ.
2. The performance and history of the machine is well documented. In 1813, it played in high-profile benefit concerts C16.N2
alongside Beethoven, Hummel, Spohr, and other Viennese musical greats. This hybrid man–machine concert, which also
featured Beethovenʼs Wellingtons Sieg Op. 91 (1813), was so profitable that Maelzel and the trumpeter travelled across
Europe, replicating the performance-spectacle with other live orchestras (Thayer 1921: 251–260).

3. Henri Decremps (1746–1826) detailed Mr van Estinʼs Wunderkammer, in which a marvellous automatical organ replicated a C16.N3
full orchestra with ʻmore precision in measure than you usually hear in instrumental performances executed by common
musiciansʼ.

4. Among its deficiencies, the reviewer notes, the trumpeter could not dynamically shape notes, which the English called C16.N4

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
ʻswellʼ, a practice that expressive musicians intuited without notated indications. From this account, it seems that the
baroque-era mezza di voce technique was retained into the early nineteenth century.

5. Thayer described the mechanics of the Panharmonicon as it originally appeared in its first home, Maelzelʼs Viennese C16.N5
Wunderkammer: ʻThe Panharmonicon combined the common instruments then employed in military bands, with a
powerful bellows—the whole thing being enclosed in a case. The motive power was automatic and the keys were touched
by pins fixed in a revolving cylinder, as in the common hand-organ or music-boxʼ.

6. Schindler describes the fraught business arrangement between Maelzel and Beethoven regarding the Panharmonicon C16.N6
composition, its orchestral arrangement, and subsequent performances. Moscheles attests that Maelzel himself conceived
and composed a good portion of the work (Schindler 1841: 148–156).

7. In Mr Gladstone, Edward Walter Hamilton commented upon a characterʼs oratorical skill similarly: ʻThe pace at which he C16.N7
spoke was an even one. He could have spoken to a metronome, though he had one pace for the House of Commons and
another pace for the platform. There was never a pause for want of an expressionʼ.

8. Gottfried Weber believed a better, age-old tempo device was well within everyoneʼs grasp. Surmounting technological C16.N8
inconveniences ʻcan in fact be done by using, instead of Maelzelʼs machine, as can be done with entire satisfaction, merely
a simple thread pendulum, i. e. any small weight as e. g. a lead ball of any size that may be preferred, suspended by a
thread; an instrument, which everyone can manufacture for himself in two minutesʼ timeʼ.

9. Beethoven was no exception in this regard: Maelzel believed the composer also required lessons on proper metronome C16.N9
p. 333 use. While coddling the composerʼs ego, Maelzel sent Beethoven a detailed instructional chart that explicated the
process of metronomic tempo calculation. ʻIt goes without saying that I do not want to give you any instruction about itʼ,
Maelzel assures Beethoven, ʻyou know the subject as well as anyoneʼ. Apparently Beethoven cherished Maelzelʼs
instructional document to such an extent that he used the back page as a notepad and shopping list, shared with his
nephew Karl (Albrecht 1996: 137, 138 n. 5).

10. It must be recognized that, despite their renown, Maelzelʼs technological marvels—his automaton trumpeter, C16.N10
Panharmonicon, and chess-playing android—were not received as obvious consumer necessities during his lifetime either.

11. On 16 July 1816, The Times ran an advertisement, endorsed by 17 eminent musicians working in Great Britain. It stated: C16.N11
ʻThe metronome of Mr. Maelzel appears to us to accomplish all that can be desired […] The greatest composers on the
Continent having already adopted the metronomical scale in indicating the time of their movements, the musical
commonwealth may, at length, expect to have a universal standard measure of musical time throughout Europeʼ.

12. On 14 February 1818, the Wiener AmZ published the inventorʼs promotional announcement, with the names of Beethoven C16.N12
and Salieri appearing as signatories: ʻMaelzelʼs Metronome is here! The usefulness of his invention will prove itself;
furthermore, all the composers of Germany, England, and France have endorsed itʼ.

13. François-Joseph Fétis (1784–1871), although uniquely approving of Maelzelʼs automated machine for conservatory C16.N13
training, nevertheless admitted, ʻChronometres of the [simple pendulum] kind […] have the double advantage of being
simple in their construction, and of trifling expenseʼ (Fétis 1842: 37). An important proponent of the French conservatory
system, Fétis thought simple pendulums ʻhave the inconvenience of not making sensible to the ear the tick or stroke of
timeʼ. Nevertheless he describes Maelzelʼs technology as a variable-length tempo pendulum: ʻEvery gradation of
movement, from slowest to the most rapid, is there expressed and represented by vibrations of the balance […] we are
[therefore] able to substitute a short rod for a very long one, and to produce great varieties of movement, by very slight
changes in the position of the central pointʼ.
14. What were described as ʻmetronomizedʼ or ʻmetronomedʼ publications—the overlaying of MM numbers upon the pre- C16.N14
existing works of renowned, o en deceased composers—became a new job for the professional editor. Metronoming
music was part of a larger editorial trend guiding the lay musician away from grossly misinterpreting a new purchase.
Sometimes in conjunction with revised rhythmic notations, publishers o ered these MM additions as value-added
features in the latest music publication. For Moschelesʼ personal justifications regarding the metronoming of Beethoven
editions, see Schindler (1841: 106–107n.).

15. Wagner also found futility in the mechanical-temporal calculation of his own compositions, and gave up any attempt to C16.N15
quantify musical time via clockwork controls. ʻIn my earlier operas I gave detailed directions as to the tempi, and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
indicated them (as I thought) accurately, by means of the Metronomeʼ, he explained in Über das Dirigiren, ʻ[but] in my later
works I omitted the metronome and merely described the main tempi in general terms, paying, however, particular
attention to the various modifications of tempo [using words]ʼ (Wagner 1897: 20–21).

16. Among many similar comments, this one from 1835 succinctly states, ʻThe too frequent use of it will make you feel ill at C16.N16
ease when you have to play without itʼ (Musical instruction 1835: 86).

17. The observation continued into the next century. ʻThe most mechanical playing imaginable can proceed from those who C16.N17
p. 334 make themselves slaves to this little musical clock, which was never intended to stand like a ruler over every minute of
the studentʼs practice timeʼ, said Serge Rachmanino (1873–1943) (Cooke 1917: 213).

18. Into the next century, some believed that the damaging, anti-musical e ects of the automatic metronome, when used C16.N18
cautiously and intelligently, could be mitigated. ʻThe metronome itself must not be used “with closed eyes”, as we should
say it in Russiaʼ, Rachmanino warned in 1913, ʻThe player must use discretionʼ.

19. ʻWhen one sees a gentleman in the rural districts, swelling with importance because he is shaking the stick in question, C16.N19
and determined to get six entire shakes into each measure of a 6/8 presto movement or die, we can but recall the term
applied to these ague-conductors in Europe. They call them “Metronomes!” ʼ (Elson 1897: 299–300).

20. In 1897, the New York Times critic heard an odd performance of Brahmsʼs Symphony No. 1 in C Minor Op. 68 (1876), where C16.N20
in the first movement, ʻThere was no rubato at all, and much of the significance of the beautiful phrasing was quite lost
because Mr. Damrosch drove his orchestra ahead with the angular rigidity of a metronomeʼ.

21. In 1891 a British Times critic noted that pianist Bernard Stavenhagen (1862–1914) performed with a strange ʻmetronomic C16.N21
accuracy of rhythmʼ that was entirely ʻill-suitedʼ to Beethovenʼs music.

22. In the pro-scientific periodical Music, Emil Liebling criticizes Paderewski for his rhythmical practices. Liebling C16.N22
correspondingly reveals his own metronomic aesthetics: ʻPerhaps the most reprehensible liberties are taken with rhythms
and tempos. In this regard Paderewski has much to answer for. Without desiring to apply the metronome to his
performances, it yet is undeniable that he hardly ever preserves the [metronomic] rhythm of a movement long enough to
give the listener an absolute idea of [metronomic] timeʼ (Liebling 1891–2: 585).

23. E.g. blind faith to a metronomic musical education is eloquently preached in the Apr. 1895 issue of the Christian Science C16.N23
Journal. The essayist, a typical middle-class mother, relates the impact of the new clockwork metronome upon her
daughterʼs daily practice: ʻNow there could be no question whether her music were right or wrong. If it [her childʼs
performance] agreed with the monitor [the metronome] it was right: if it disagreed it was wrong and there could be also no
o enseʼ (H.L.B. 1895: 375–376).

24. She continues: ʻso many notes in a trill, in scale, in arpeggio, octaves, chords, etc. (which implies a hand-position and C16.N24
good fingers). It was a great boon to piano-teachers and students when this minute system was condensed and presented
to the public, simplifying as well as making thorough this one branch of pianistic teachingʼ (Chittenden 1907: 145).

25. Brahms concluded: ʻI am of the opinion that metronome marks go for nothing. As far as I know, all composers have as yet C16.N25
retracted their metronome marks in later years. Those figures which can be found before some of my compositions—good
friends have talked them into me; for I myself have never believed that my blood and a mechanical instrument go very
well togetherʼ. Also in Vechten (1920: 218).

26. ʻ304. How to memorize quarter note=60. / Guided by the second-hand of a watch, count the seconds as follows:/ One- C16.N26
thousand one, one-thousand two, one-thousand three, etc./ With a little practice the speed of this rhythm can be readily
recalled when the / accented syllables, one, two, three, etc., will indicate the value of each quarter-note in the standard
tempo quarter note=60. / From this, other tempi may be at least approximated. / 305. To discover the rate of speed of a
composition during its performance count the number of beats in a quarter of a minute and multiply by fourʼ.

p. 335 27. Without further discussion of the sensation or perception of internal pulse and movement, the pedagogue—maintaining a C16.N27
semblance of musical-temporal traditionalism—finally states: ʻRemember, when you learn to play, use the metronome
only a moment, to show you what time was in the composerʼs mind. Never depend on it wholly to keep steady timeʼ.

28. He continues: ʻThe slow trill is to be practiced continually, with the metronome mark raised in later lessons until 96 is C16.N28
reached […] But speed should never be increased unless the movements are kept correctʼ.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
29. The inventor-educator Frederick Franz also considered his black-box metronome the right practice tool for repertoire by C. C16.N29
P. E. Bach and Chopin, for ʻimproved rhythmic performanceʼ (Franz 1964: 29–31).

30. With its automatic sound predominating in modern pedagogies and metronome-reliant compositional practices, the C16.N30
machineʼs visual aspect (the pendulum) was o en seen as either a mere ornament or noticeable distraction. One
professional solved the problem of the metronomeʼs now-distracting pendulum thus: ʻI found that by placing the
objectionable machine on the floor at my feet, I am not distracted by the motion of the arm of the metronome and Iʼm not
tempted to look at the instrumentʼ. The comment implies that the constant sound of the metronome was not a distraction
while playing. Franz credits ʻN. G. Abbott, Etude, October, 1923ʼ.

31. Beyond establishing mathematical-mechanical precision as the theoretical basis for musical temporality, modern music C16.N31
training o en required the playing of actual compositions to metronomic sound—the great sin of historical musicality:
ʻTime-beating exercises may be made by using short pieces of moderate di iculty and tapping the time value of the notes
to the beat of the metronome. Easy movements from the Mozart sonatas will furnish abundant materialʼ (Brower 1920:
18).

32. ʻRhythm is becoming constricted into ever-narrower meanings; it is already used almost as a synonymous term for the C16.N32
regularity of the audible beatʼ, claimed New York Times critic John Martin in 1929.

33. ʻWe are meeting a tendency in music to discard the barʼ, observed Arthur L. Salmon in Musical Times (1920), ʻThe definite C16.N33
abandonment has not yet been fully brought into practice, yet the [sensed] bar as a controller of measure is being more
and more slightedʼ. A new metronomic modality was in operation, Salmon implied, as ʻ[t]he old accentual teaching is
largely discounted—that teaching which impressed on a beginner the necessity for emphasizing every first and third beat
in common time, for instanceʼ.

34. Franz credits ʻEugene F. Markes, Etude, June, 1925ʼ. C16.N34

35. Some influential turn-of-the-century pedagogies urged children and adults to behave more scientifically, automatically, C16.N35
and rhythmically normal—more like Maelzelʼs forgotten automata from the past century. ʻNever again [in a childʼs
formative years] will there be such susceptibility to drill and discipline, such plasticity to habituation, or such ready
adjustment to new conditionsʼ, the dean of modern America education G. Stanley Hall (1846–1924) proclaimed in 1904,
adding, ʻIt is the age of external and mechanical training. Reading, writing, drawing, manual training, musical technic […]
have now their golden hourʼ. This text is a reduction of Hallʼs 1904 book Adolescence.

36. The theories of George Antheil, in many ways Stravinskyʼs rhythmical acolyte, are indicative of Machine Age musical- C16.N36
temporal trends. Antheil championed his proprietary music-notation plan: ʻAny system denoting the [musical] movement
p. 336 of time in space in an unmathematical, sentimental, and literary way is to be condemned as incomprehensible. / One
can only measure spaces with mathematics!ʼ (Antheil n.d.: 3–4). Antheilʼs failed, calculative notational system was
preceded most notably and successfully by Dalcrozeʼs Eurythmics.

37. The modernist Herman Scherchen supported a similarly uncritical conviction in metronomic time, as he sought to C16.N37
perform Bach scientifically: ʻThe metronome marking which suits both forms of [Bachʼs chorale melody] is: printed
quarter / eighth notes=72 (Maelzelʼs metronome). This allows listeners to think in terms of two quavers together […] this
tempo […] is the human normal value, by which seventy two pulse beats and eighteen full respirations are produced a
minuteʼ (Scherchen 1950: 87–88).

38. Johann Maelzel reportedly knew the di erence between the ideal human performance and the most perfect clockwork C16.N38
tempo-machine: ʻWe do not contend that an accomplished musician should perform an entire piece according to the
metronome; for all expression would be paralysed by such enslavementʼ. Originally from ʻInformation on the Metronome
by J. Maelzelʼ (May 1818), housed in the Library of the Gesellscha der Musikfreunde, Vienna.

39. Schindler recounts Beethovenʼs troubles when metronoming his compositions for publishers. Schindler claimed ʻ[during C16.N39
the winter of 1825–26] in my presence [Beethoven] ascertained that the metronomic signs in the printed scores [of the
symphonies] were faulty, in fixing the tempi too quick; and, indeed, he declared that many of those metronomic signs
were not authorized by himʼ (1841: 104–105, 116). Schindler also recalled that Beethoven occasionally marked the same
music with di erent metronomic indications (pp. 116–117).

40. Herbert von Karajan seemed to endorse the mechanized beat-training systems integral to modern musicianship: ʻAnd if C16.N40

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
pupils are not taught the basic disciplines of rhythm—then things become impossibleʼ (Osborne 1989: 96).

41. ʻBy setting the metronome at a speed you feel is comfortable for your successful execution of the passage, you have C16.N41
eliminated the necessity for [self-] governing the speed of your playing—the metronome will do it for youʼ, instructed this
unexceptional piano method (Miller 2007: 214–215).

42. Pedagogies such as this one o en reiterate the now normative procedures for dispassionately entraining musicians into C16.N42
rhythmic automatism: ʻOnce you have learned a new piece of music, prove to yourself that you know it well by playing it
with the metronome tickingʼ (Miller 2007: 214–215).

43. Unlike many musicians educated through metronome-reliant pedagogies of the 20th c., Beethoven did not grow up C16.N43
practising to a clockwork metronome.

44. Segment at approximately 7 minutes, 30 seconds (Radiolab 2013). C16.N44

45. Dialogue begins at approximately 13 minutes (ibid.). C16.N45

46. Popular pedagogies, including the Suzuki Method, o en apply the metronomic click track as an a priori condition for C16.N46
accurate rhythmical movement. Future research will reveal these processes more fully.

47. In 1917 Constantin von Sternberg noted the technological commonality: ʻWe must regard the inventor a little closer and C16.N47
consider the time in which the misfortune of his invention [the metronome] happened […] He constructed all kinds of
mechanical instruments, such as an automaton trumpeter, a mechanical orchestra, a “panharmonium”, I believe also a
mechanical chess-player and the metronome!ʼ (Stenberg 1917: 79–80).

48. When someone blindly submitted their innate rhythmical sense of pulse and movement to metronomic sound, Sternberg C16.N48
said, ʻevery distinction between the living musician and the mechanical self-playing machine practically vanishesʼ (1917:
79–80).
p. 337
References C16.S7

Ackermann, W. (c.1812). Description of Mr. Ackermannʼs musical regulators or pendulums, and directions for using them. Harrison C16.P131
& Rutter.

Adams, E. (1903). Electric metronome. U.S. Patent No. 734,032 (filed 30 July 1902). C16.P132
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Advertisement (1897). Music 11: inside cover. C16.P133
Google Scholar Google Preview WorldCat COPAC

Albrecht, T. (1996). Letters to Beethoven and other correspondence, vol. 2. University of Nebraska Press. C16.P134
Google Scholar Google Preview WorldCat COPAC

Amar, J. (1918). The physiology of industrial organization and the re-employment of the disabled. Library Press. C16.P135
Google Scholar Google Preview WorldCat COPAC

An historical miscellany of the curiosities and rarities in nature and art. Comprising new and entertaining descriptions of the most C16.P136
surprising volcanos, caverns, cataracts, whirlpools, […] in every part of the habitable world (1794–1800), vol. 3. London.

Antheil, G. (n.d.). Antheilicized notation, or music for everybody who can tell one from two and two from three, without previous C16.P137
experience anyone can now read music at sight. Instantaneous system of reading music. Antheil Collection. Library of Congress,
Box 10.
Google Scholar Google Preview WorldCat COPAC

Aristotle (1959). Nichomachean ethics. Oxford University Press. C16.P138


Google Scholar Google Preview WorldCat COPAC

Arrington, J. E., and Ohl, J. F. (1960). John Maelzel, master showman of automata and panoramas. Pennsylvania Magazine of C16.P139
History and Biography 84: 56–92.
Google Scholar WorldCat

Attwood, T. (1816). To the Editor. The Times, 16 July, 3. C16.P140


Google Scholar WorldCat

B., C. A. (1888). George Groveʼs analyses of Beethoven. Musical World, vol. 850. C16.P141
WorldCat

B., H. L. (1895). A lesson from the metronome. Christian Science Journal 13: 375–377. C16.P142
WorldCat

Barnum, P. T. (1855). The autobiography of P. T. Barnum. Ward & Lock. C16.P143

Bartky, I. (2000). Selling the true time: Nineteenth century timekeeping in America. Stanford University Press. C16.P144
Google Scholar Google Preview WorldCat COPAC

Beethoven, L. von, and Salieri, A. (1818). Erklarung. Wiener AmZ 2(7): 58f. C16.P145
Google Scholar WorldCat

Berhard, K. (1828). Travels through North America, during the years 1825 and 1826, vol. 1. Carey, Lea & Carey. C16.P146
Google Scholar Google Preview WorldCat COPAC

Berlioz, H. (1856). A treatise on modern instrumentation and orchestration. Novello, Ewer, & Co. C16.P147
Google Scholar Google Preview WorldCat COPAC
Bingham, T. (2012). The Tony Bingham metronome collection: An illustrated checklist. Tony Bingham. C16.P148
Google Scholar Google Preview WorldCat COPAC

Bonus, A. (2013). A timely musical discourse, or a music treatise from lost times, pt 1. Current Musicology 95: 319–347. C16.P149
Google Scholar WorldCat

Bonus, A. (2014). Metronome. Oxford Handbooks Online. Oxford University Press. http://www.oxfordhandbooks.com C16.P150
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Bonus, A. (2017). Refashioning rhythm: Hearing, acting, and reacting to metronomic sound in the experimental sciences, c.1875– C16.P151
1920. In I. Biddle and K. Gibson (eds), Cultural histories of noise, sound and listening in Europe, 1300–1918, 76–105. Routledge.
Google Scholar Google Preview WorldCat COPAC

Bowman, E. M. (1910). Letters from a musician to his nephew. In The Essentials of Music, vol. 1. Irving Squire. C16.P152
Google Scholar Google Preview WorldCat COPAC

Brower, H. M. (1917). Piano mastery: Second series. Frederick A. Stokes. C16.P153


Google Scholar Google Preview WorldCat COPAC

Brower, H. M. (1920). Self-help in piano study. Frederick A. Stokes. C16.P154


Google Scholar Google Preview WorldCat COPAC

Brown, C. (1999). Classical and Romantic performing practice, 1750–1900. Oxford University Press. C16.P155
Google Scholar Google Preview WorldCat COPAC

Buchner, A. [n.d.]. Mechanical musical instruments. Batchworth Press. C16.P156


Google Scholar Google Preview WorldCat COPAC

Chittenden, K. S. (1907). Report of piano conference. Studies in Musical Education History and Aesthetics: Papers and Proceedings C16.P157
of the Music Teachersʼ National Association (27–31 December): 144–151.
WorldCat

p. 338 Christiani, A. F. (1885). The principles of expression in pianoforte playing. Harper & Brothers. C16.P158
Google Scholar Google Preview WorldCat COPAC

Constant (1907). Recollections of the private life of Napoleon, vol. 2. Saalfield. C16.P159
Google Scholar Google Preview WorldCat COPAC

Cooke, J. F. (1917). Great pianists on piano playing: Study talks with foremost virtuosos. Theodore Presser. C16.P160
Google Scholar Google Preview WorldCat COPAC

Czerny, C. (1839). Complete theoretical and practical piano forte school, vol. 3. R. Cocks. C16.P161
Google Scholar Google Preview WorldCat COPAC

Decremps, H. (1785). The conjurer unmasked; Or, la magie blanche dévoilée: Being a clear and full explanation of all the surprizing C16.P162
performances exhibited. Printed for C. Stalker.
Google Scholar Google Preview WorldCat COPAC

Delafield, J. (1857). Mysticism and its results: Being an inquiry into the uses and abuses of secrecy. Edwards & Bushnell. C16.P163
Google Scholar Google Preview WorldCat COPAC

Descartes, R. (1850[1637]). Discourse on the method of rightly conducting the reason, and seeking the truth in the sciences. C16.P164
Sutherland & Knox.
Google Scholar Google Preview WorldCat COPAC
Elson, A. (1918). Modern piano methods. In L. C. Elson (ed.), Modern Music and Musicians, vol. 1, 69–82. University Society. C16.P165
Google Scholar Google Preview WorldCat COPAC

Elson, L. C. (1897). The realm of music. New England Conservatory. C16.P166


Google Scholar Google Preview WorldCat COPAC

Encyclopedia Britannica (1884). The Encyclopedia Britannica. A Dictionary of Arts, Sciences, and General Literature, vol. 16 (9th C16.P167
edn). J. M. Stoddart.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Fétis, F. J. (1842). Music explained to the world: Or, how to understand music and enjoy its performance. Benjamin Perkins. C16.P168
Google Scholar Google Preview WorldCat COPAC

Finck, H. T. (1909). Success in music and how it is won. Scribnerʼs. C16.P169


Google Scholar Google Preview WorldCat COPAC

Fiske, D. W. (1859). The book of the First American Chess Congress. Sampson Low. C16.P170
Google Scholar Google Preview WorldCat COPAC

Franz, F. (1964). Metronome techniques. Frederick Franz. C16.P171


Google Scholar Google Preview WorldCat COPAC

Gardinerʼs music of nature (1838). New York Review 5: 46–67. C16.P172


Google Scholar Google Preview WorldCat COPAC

Goudsmit, M. J. (1930). Music time indicator. U.S. Patent #1,783,537. Filed 2 Aug. 1923. C16.P173
Google Scholar Google Preview WorldCat COPAC

Gould, H. F. (1829). To the automaton chess player. In Landmark anthologies: Specimens of American poetry. S. G. Goodrich. C16.P174
http://ebooks.ohiolink.edu
Google Scholar Google Preview WorldCat COPAC

Grew, S. (1922). The art of the player-piano. Kegan Paul. C16.P175


Google Scholar Google Preview WorldCat COPAC

Grove, G. (1880–1889). Maelzel. In A dictionary of music and musicians, vol. 2. Macmillan. C16.P176
Google Scholar Google Preview WorldCat COPAC

Grund, F. J. (1837). The Americans in their moral, social, and political relations. Marsh, Capen, & Lyon. C16.P177
Google Scholar Google Preview WorldCat COPAC

Hall, G. S. (1907). Youth: Its education, regimen, and hygiene. D. Appleton. C16.P178
Google Scholar Google Preview WorldCat COPAC

Hamilton, C. G. (1916). How to use the metronome correctly. Theodore Presser. C16.P179
Google Scholar Google Preview WorldCat COPAC

Hamilton, E. W. (1898). Mr. Gladstone. Scribnerʼs. C16.P180

Hanchette, H. G. (1905). Art of the musician. Macmillan. C16.P181


Google Scholar Google Preview WorldCat COPAC

Herr Stavenhagenʼs recital (1891). The Times, 25 Nov., 3. C16.P182


Google Scholar Google Preview WorldCat COPAC
Ho mann, E. T. A. (1908). Automata. In Serapion Brethren, vol. 1. George Bell. C16.P183
Google Scholar Google Preview WorldCat COPAC

Holcro , T. (1798). Knave; or not? Printed for G. G. & J. Robinson. C16.P184


Google Scholar Google Preview WorldCat COPAC

Holcro , T. (1801). Deaf and dumb, or the Abbé de lʼÉpée. Samuel French. C16.P185
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Holland, B. (2008, February 6). When histrionics undermine the music and the pianist. New York Times. C16.P186
http://www.nytimes.com/2008/02/06/arts/music/06look.html
WorldCat

Holmes, O. W. (1866[1863]). The human wheel, its spokes and felloes. In Soundings from the Atlantic, 282–327. Tickner & Fields. C16.P187
Google Scholar Google Preview WorldCat COPAC

Hopkins, A. (1898). Magic: Stage illusions and scientific diversions. Munn & Co. C16.P188
Google Scholar Google Preview WorldCat COPAC

Kennedy, J. P. (1850). Memoirs of the life of William Wirt II. Lea & Blanchard. C16.P189
Google Scholar Google Preview WorldCat COPAC

Kotzschmar, M. A. T. (1907). Half-hour lessons in music. Oliver Ditson. C16.P190


Google Scholar Google Preview WorldCat COPAC

Krehbiel, H. E. (1898). Music and manners in the Classical period: Essays. Scribnerʼs. C16.P191
Google Scholar Google Preview WorldCat COPAC

Kullak, F. (1901). Beethovenʼs piano-playing. G. Schirmer. C16.P192


Google Scholar Google Preview WorldCat COPAC

p. 339 Lamb, J. (2014). A matter of time: The science of rhythm and the groove. Swing Press. C16.P193
Google Scholar Google Preview WorldCat COPAC

Landes, D. S. (2000). Revolution in time: Clocks and the making of the modern world. Harvard University Press. C16.P194
Google Scholar Google Preview WorldCat COPAC

Langenus, G. (1933). Rhythm-Builder. Ensemble Music Press. C16.P195


Google Scholar Google Preview WorldCat COPAC

Lawrence, V. B. (1988). Strong on music, vol. 1. Oxford University Press. C16.P196


Google Scholar Google Preview WorldCat COPAC

Liebling, E. (1891–2). A pianistic retrospect. Music 1: 583–586. C16.P197


WorldCat

Listening to music (1897, March 6). New York Times, 6 Mar., RB2. C16.P198
WorldCat

Liszt, F. (1894). Letters of Franz Liszt II. Scribnerʼs. C16.P199


Google Scholar Google Preview WorldCat COPAC

Maelzel, John (1837). The American Magazine of Useful and Entertaining Knowledge, 1 Feb., 196–7. C16.P200
Google Scholar Google Preview WorldCat COPAC
Marshall, E. (1901). Scientific child study. Frank Leslieʼs Popular Monthly 51: 419–430. C16.P201
https://search.proquest.com/americanperiodicals/
Google Scholar WorldCat

Martin, J. (1929). The dance: Listening in. New York Times, 31 Mar., 108. C16.P202
Google Scholar WorldCat

Marx, A. B. (1853). Universal school of music. Robert Cocks. C16.P203


Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Marx, A. B. (1895). Introduction to the interpretation of Beethoven piano works. Clayton F. Summy. C16.P204
Google Scholar Google Preview WorldCat COPAC

Mattheson, J. (1737). Kern melodischer Wissenscha . Hamburg. C16.P205


Google Scholar Google Preview WorldCat COPAC

McLaughlin, J. M. (1902). Elements and notation of music. Ginn. C16.P206


Google Scholar Google Preview WorldCat COPAC

Metzner, P. (1998). Crescendo of the virtuoso. University of California Press. C16.P207


Google Scholar Google Preview WorldCat COPAC

Miller, N. (2007). The piano lessons book: The piano studentʼs guide for getting the most out of practicing, lessons, your teacher and C16.P208
yourself. CreateSpace.
Google Scholar Google Preview WorldCat COPAC

Minshull, J. (1803). A comedy, entitled the sprightly widow. Printed for the Author. C16.P209
Google Scholar Google Preview WorldCat COPAC

Music (1856). The Albion, a Journal of News, Politics, and Literature, 8 Mar., 115. C16.P210
Google Scholar Google Preview WorldCat COPAC

Musical instruction (1835). The Zodiac, a Monthly Periodical, Devoted to Science, Literature, and the Art (Dec.): 86. C16.P211

Musical review (1846). Harbinger, Devoted to Social and Political Progress, 7 Mar., 204. C16.P212
Google Scholar Google Preview WorldCat COPAC

Musical review (1847). Harbinger, Devoted to Social and Political Progress, 27 Feb., 185. C16.P213
Google Scholar Google Preview WorldCat COPAC

Musser, C. O. (1966). Electric metronome. U.S. Patent #3,263,551. Filed 7 May 1965. C16.P214
Google Scholar Google Preview WorldCat COPAC

Osborne, R. (1989). Conversations with Karajan. Oxford University Press. C16.P215


Google Scholar Google Preview WorldCat COPAC

Panharmonicon (1825). Western Recorder, 23 Aug., 136. http://www.proquest.com/ C16.P216


Google Scholar Google Preview WorldCat COPAC

Parker, J. (1825). Musical biography. Stone & Fovell. C16.P217


Google Scholar Google Preview WorldCat COPAC

Parsons, A. R. (1886). The science of pianoforte practice. G. Schirmer. C16.P218


Google Scholar Google Preview WorldCat COPAC
Philharmonicus (1817). An explanation of the notation employed in the scale of Maelzelʼs metronome. Edinburgh Magazine and C16.P219
Literary Miscellany 1: 223–224.
WorldCat

Philip, R. (2004). Performing music in the age of recording. Yale University Press. C16.P220
Google Scholar Google Preview WorldCat COPAC

Poe, E. A. (1836). Maelzelʼs chess player. Southern Literary Messenger 2: 318–326. C16.P221
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Quarterly Musical Magazine and Review (1821). Vol. 3, p. 302. C16.P222

Radiolab (2013). Speedy beet [radio segment]. The Power of Music (radio episode). New York: WNYC Radio. C16.P223
http://www.radiolab.org/story/269783-speedy-beet/

Register of Debates in Congress XIV (1837). Gales & Seaton. C16.P224


Google Scholar Google Preview WorldCat COPAC

Rothschild, F. (1961). Musical performance in the times of Mozart and Beethoven. Oxford University Press. C16.P225
Google Scholar Google Preview WorldCat COPAC

Sachs, C. (1953). Rhythm and tempo. Norton. C16.P226


Google Scholar Google Preview WorldCat COPAC

Salmon, A. L. (1920). Shall we retain the bar? Musical Times 61: 587. C16.P227
Google Scholar WorldCat

Scherchen, H. (1950). The nature of music. Denis Dobson. [Vom Wesen der Musik, 1946.] C16.P228
Google Scholar Google Preview WorldCat COPAC

Schillinger, J. (1976). Encyclopedia of rhythms instrumental forms of harmony. Da Capo Press. C16.P229
Google Scholar Google Preview WorldCat COPAC

p. 340 Schindler, A. (1841). The life of Beethoven: Including his correspondence with his friends, numerous characteristic traits, and C16.P230
remarks on his musical works vol. 2, ed. I. Moscheles. Henry Colburn.
Google Scholar Google Preview WorldCat COPAC

Scripture, E. W. (1895). Thinking, feeling, doing. Chautauqua-Century Press. C16.P231


Google Scholar Google Preview WorldCat COPAC

Seashore, C. E. (1915). The measurement of musical talent. Musical Quarterly 1: 129–148. C16.P232
Google Scholar WorldCat

Seashore, C. E. (1919). The psychology of musical talent. Silver/Burdett. C16.P233


Google Scholar Google Preview WorldCat COPAC

Snyder, A. (2001). The rhythm reader II. Hal Leonard. C16.P234


Google Scholar Google Preview WorldCat COPAC

Spohr, L. (1833). Grand violin school. Wessel. C16.P235


Google Scholar Google Preview WorldCat COPAC

Sternberg, C. von (1917). Ethics and esthetics of piano-playing. G. Schirmer. C16.P236


Google Scholar Google Preview WorldCat COPAC
Substitute for Maelzelʼs metronome (1840). The Musical Magazine; or, Repository of Musical Science, Literature 67. C16.P237
http://www.proquest.com
Google Scholar Google Preview WorldCat COPAC

Sweetwater (2018). Soundbrenner pulse vibrating metronome. https://www.sweetwater.com/store/detail/SBPulse— C16.P238


soundbrenner-pulse-vibrating-metronome
Google Scholar Google Preview WorldCat COPAC

Thayer, A. W. (1921). The life of Ludwig van Beethoven, vol. 2. Beethoven Association. C16.P239

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

The metronome (1836). Musical Magazine, Apr., 304. C16.P240


Google Scholar Google Preview WorldCat COPAC

The Symphony Society (1898). The New York Times, 23 Jan., 16. C16.P241
Google Scholar Google Preview WorldCat COPAC

The violin. Hogarthʼs musical history (1839). Musical Magazine, 17 Aug., 270. C16.P242
Google Scholar Google Preview WorldCat COPAC

Thomson, T., Sandford D. K., and Cunningham A. (1836). The popular encyclopedia, vol. 1. Blackie & Son. C16.P243
Google Scholar Google Preview WorldCat COPAC

Timbs, J. (1856). A year-book of facts in science and art. David Bogue. C16.P244
Google Scholar Google Preview WorldCat COPAC

To the Editor (1816). The Times, 16 July, 3. C16.P245


Google Scholar Google Preview WorldCat COPAC

Ure, A. (1835). The philosophy of manufactures: Or an exposition of the scientific, moral, and commercial economy of the factory C16.P246
system of Great Britain. Charles Knight.
Google Scholar Google Preview WorldCat COPAC

Vechten, C. V. (1920). In the garret. Knopf. C16.P247


Google Scholar Google Preview WorldCat COPAC

Vorkapich, S. (director) (1943). Conquer by the clock (film). RKO-Pathé. Retrieved from C16.P248
https://archive.org/details/0331_Conquer_by_the_Clock_09_00_55_00
Google Scholar Google Preview WorldCat COPAC

Wagner, R. (1897). On conducting [Über das Dirigire]. William Reeves. C16.P249


Google Scholar Google Preview WorldCat COPAC

Weber, C. M. von. (1827). Letter by C. M. Von Weber on the performance of dramatic song. Harmonicon 5(1): 219–220. C16.P250
Google Scholar WorldCat

Weber, G. (1841). General music teacher: Adapted to self-instruction, both for teachers and learners; Embracing also an extensive C16.P251
dictionary of musical terms. J. H. Wilkins & R. B. Carter.
Google Scholar Google Preview WorldCat COPAC

Webster, D. (1903). Technical progress and prosperity. In The Writings and Speeches of Daniel Webster Hitherto Uncollected, vol. 1, C16.P252
63–78. Little, Brown.
Google Scholar Google Preview WorldCat COPAC

Whitrow, G. J. (1989). Time in history: Views of time from prehistory to the present day. Oxford University Press. C16.P253
Google Scholar Google Preview WorldCat COPAC

Willis, R. S. (1853). On musical tempo. Musical World and New York Musical Times, 11 June, 82. C16.P254
WorldCat

Wonderful toys (1852). The North American Miscellany and Dollar Magazine, 1 Feb., 103. C16.P255
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470256 by Shanghai Jiao Tong University user on 26 May 2023
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
CHAPTER

17 Rhythm Quantization: Notes


C17 on the History of a
Technocultural Practice 
Landon Morrison

https://doi.org/10.1093/oxfordhb/9780190947279.013.18 Pages 341–C17.P122


Published: 08 December 2021

Abstract
This chapter sketches a general history of rhythm quantization as a widespread practice in popular
music culture. Quantization—a sound technology that automatically maps microrhythmic uctuations
onto the nearest beat available within a prede ned metric grid—challenges traditional notions of
musicking as an embodied activity that is grounded in the co-presence of human agents. At the same
time, it encapsulates cultural and cognitive processes that are entirely human, tting into a broader
historical shift towards chronometric precision in Western music. Questions arising from this apparent
contradiction are taken up in this chapter, which situates rhythm quantization as an emergent
technocultural practice, examining its attendant technologies and requisite structures of music-
theoretical knowledge, as well as its reception within the context of di erent musical genres.

Keywords: rhythm quantization, microrhythm, metric grid, technoculture, sound technology, popular
music
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Introduction C17.S1

FUELLED by the rise of digital sound technologies, rhythm quantization has become an increasingly C17.P1
widespread practice in popular music, yet there has been little re ection on how it alters people’s
experiences of musical time. Out of sight and out of mind, quantization operates at the level of silent
ubiquity, slipping undetected into daily life. As Grammy-winning producer Glen Ballard recently put it:

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
‘almost everything people hear now, certainly in America, is quantized. And so, if you put a real drummer
1
on, sometimes people don’t get that. They need that extra little […] ‘authority’ of the regular time.’ If this
assessment is correct, then it raises a number of important questions: How did the ‘authority’ of rhythm
quantization become so entrenched, so naturalized and self-evident, as to be nearly inaudible to modern
ears? What were the attendant technologies and requisite structures of knowledge that facilitated its rise as
a technocultural practice? And more broadly, how does it in ect the social and aesthetic dynamics of
popular music culture?

In this chapter, I will consider each of these questions; but to begin, it may be helpful to clarify what exactly C17.P2
is meant by ‘rhythm quantization’. Fundamentally, it refers to the process of mapping the small,
microrhythmic uctuations found in human performance to the nearest beat available within a pre-de ned
2
metric framework. Anecdotally, anyone who has ever tried their hand at transcription will be familiar with
p. 342 this basic procedure, as the process of notating sounds involves a similar conversion of dynamic
phenomena into discrete form. In both cases, calculated approximations must be made to strike a careful
balance between precision and e ciency, detail and legibility; but whereas the e ects of notation are
con ned to a purely symbolic register, rhythm quantization hitches symbols to actions, recon guring
temporal relationships between actual sounds. In this way, quantization produces an aural manifestation of
what notation only signi es, suturing sign to sound in a way that makes digital representations of rhythm
interactive, operable, and programmable.

To demonstrate how quantization works, Figures 17.1a–d present multiple instances of a simple, disco- C17.P3
inspired ‘four-on-the- oor’ drum beat. The rst example (17.1a) shows the two-bar pattern in
conventional notation, where the low–mid–high placement of notes on the sta is indicative of a kick
drum, snare, and hi-hat cymbals, respectively. In the next two examples, Ableton Live software was used to
program a MIDI version of the same pattern, shown here in a before–after sequence as originally performed
3
(17.1b) and after being quantized (17.1c). The last example (17.1d) shows an audio recording of the pattern
that has been quantized using digital editing tools.

Pictured in Figures 17.1b and 17.1c, the MIDI sequencer in Live is rendered skeuomorphically as a virtual C17.P4
piano roll, where each horizontal row represents a di erent drum sound; in this case, the lowest row is a
4
kick drum, the middle two rows are hi-hat cymbals (closed and open), and the top row is for snare drum.
Vertical lines in the background mark sixteenth subdivisions, which have been designated as the ‘quantum
value’ (i.e. fastest-moving pulse layer) of the entire metric structure. Looking closely at Figure 17.1b, there
are slight imperfections in the rhythm, detectable in those places where rectangles misalign with the
underlying grid. Setting aside the prospect that an expert drummer might have performed with better
timing, these incongruences point to a larger musical fact: no matter how accomplished the performer, it is
impossible to play with absolute metric precision. There will always be a certain degree of microrhythmic
uctuation, a bit of natural jitter in any human performance. And that is where quantization enters the
equation: shown in Figure 17.1c, the quantized representation of the same pattern exhibits perfect metric
uniformity, such that the onset of each rhythmic duration aligns with the nearest sixteenth note in the bar.
All performative deviations from clock-time, whether intentional or not, have been rmly xed within the
squared-o con nes of a referential grid.
More recently, it has become possible to quantize audio recordings of live performances. Figure 17.1d shows C17.P5
a recording of the original four-on-the- oor pattern, which was quantized using Live’s ‘time-warping’
feature. Because the content of an audio le is not pre-packaged as discrete bits of information, the
quantization process here is slightly more complicated than in the case of MIDI. First, the software
automatically detects transients in the waveform, searching for brief, high-amplitude spikes in the
spectrum that signal the onset of a sound. Then, presuming these transients generally align with the
music’s underlying pulse, the software tags them with beat markers and moves them onto the grid,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
stretching or compressing the surrounding audio proportionally to ll any resultant gaps in time. Options
p. 343 for automatically quantizing audio have been available for over a decade, but they can still be unreliable,
often requiring additional hands-on nesse and tweaking to ensure appropriate musicality.

Figure 17.1

ʻFour-on-the-floorʼ rhythm in di erent media representations: (a) notated, (b) programmed with MIDI in real-time, (c) MIDI C17.F1
version a er quantization, and (d) quantized audio waveform. For clarity color version of the figure 17 b,c and d are included in

the accompanying website

The preceding examples follow a basic de nition of rhythm quantization, but most digital audio C17.P6
workstations (DAWs) today, including Live, o er an array of lters allowing users to apply the technique in
more nuanced ways. For instance, it is typically possible to control how much quantization is applied, and to
which rhythmic units, by adjusting the relative strength and sensitivity of the process. With more advanced
options, users can apply swing to already quantized rhythms, or arti cially re-humanize them with
p. 344 controlled doses of randomization; in some programs, it’s even possible to clone the groove pro le of an
5
existing audio recording for later use as a customized quantization template. Taken together, these
features enable a division of musical time into an assortment of parameters that can be measured and
manipulated at the level of milliseconds, yielding musical results that range from perfect metric conformity
to ne-tuned simulations of expressive human timing.
By almost any measure, the capacity to control musical time with such precision marks a signi cant rupture C17.P7
with the past, yet quantization has received little notice from within the music research community. Indeed,
most writing on the topic appears in music technology magazines, online blogs, instructional videos and
texts geared towards practising musicians. Popular journals such as Keyboard or Sound on Sound, and more
recently online publications like Attack Magazine and Resident Advisor, routinely feature articles dealing with
quantization at some level, whether it’s in the form of practical advice (Aiken 2000; Buchanan 2014),
surveys of vintage drum machines (Wilson 2016), or interviews with producers about their approaches to

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
quantization in the studio (Houghton 2017). Likewise, standard how-to manuals for studio production
almost always feature at least one chapter on quantization techniques (Langford 2014; Pejrolo 2011). The
thrust of the writing in these forums is primarily pragmatic, relaying useful information about how to apply
quantization in popular music contexts, without bogging readers down in technical jargon or overly
complicated equations.

Outside of popular media, rhythm quantization has been a primary concern of empirical research in the C17.P8
cognitive and computer sciences, especially in the context of software for beat-induction, tempo-tracking,
6
and automatic transcription. With varying degrees of success, these programs have attempted to reconcile
expressive timing with the discrete properties of notation, a task that assumes musical time can be split into
two layers: a static background structure represented by notated metre and a uid foreground of
continuously shifting rhythms. The algorithms used to model this duality have tended to rely heavily on
7
psychoacoustics research that focuses primarily on tonal music from the Western classical tradition. As a
result, software based on these cognitive models has been predictably inscribed with an invisible set of
musical values—speci cally, values that enact a literal rationalization of rhythm into ideal time points, all
of which are exact multiples or divisions of a constant beat.

Drawing on these disparate strains of discourse, the present chapter maps a general history of rhythm C17.P9
quantization as a technology that is at once old and new, situating it in relation to a long-running aspiration
towards chronometric precision in Western culture, while at the same time recognizing its novel e ects on
popular music production. Revisiting the story of its invention, I question the widely accepted attribution of
quantization to a single source and suggest a revised account, in which the technology gradually coalesced
around a network of actors. The picture that emerges is illustrative of Gilbert Simondon’s claim that ‘what
p. 345 resides in the machines is human reality, human gesture xed and crystallized into working structures’
(2017 [1958]: 13). From this vantage point, I argue that quantization is not a mere technological apparatus,
but a kind of musical machine that encapsulates cultural knowledge, social processes, people and
institutions, and Western beliefs about the ontology of musical time.

After examining key digital artefacts associated with the inception of rhythm quantization, the chapter C17.P10
broadens in scope, reframing the technology in relation to prior methods for regulating musical time.
Historical precedents can be found in the use of click tracks and metronomes, both of which rationalized
tempo into a quanti able series of discrete pulses. These earlier time-keeping tools paved the way for
quantization by bending the performer’s will towards an external referent for absolute temporal objectivity.
They also codi ed in material form an understanding of rhythm and metre in Western music that can be
traced back to at least the eighteenth century. Hence, we can speak of a vital phase of what Lewis Mumford
calls ‘cultural preparation’ (1934: 9), without which the very notion of measuring musical time in such a
manner would have been inconceivable. Following the dual paths of objects and ideas, the descriptive
account presented below o ers a longue durée perspective on the origins of rhythm quantization, showing
how the practice intersects with debates about the intermingling of subjective and objective temporalities in
diverse musical contexts.

The chapter concludes by returning to the present moment, surveying the reception of quantization across a C17.P11
broad spectrum of musical genres and considering its cultural impact in light of its global reach. Paratextual
analysis of public statements by musicians, producers, and critics provides insight into why some genres
have fully embraced the technology while others have sought distance from its mechanizing e ects. Casting
a wide net, the closing section continues to push into new terrain, orienting the reader within a sprawling
eld of ongoing debates from across an array of styles and disciplines. Although it is impossible to give an
exhaustive account of such an expansive subject, the capsule history outlined here nevertheless endeavors
to o er a panoramic view of how rhythm quantization emerged as a vibrant technocultural practice, and
how it continues to enable new modes of musical expression, blurring boundaries between machines and
bodies by challenging dichotomous assumptions that often lie at the heart of popular music culture.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Stumbling onto the Grid? The Story of Roger Linnʼs LM-1 Drum C17.S2

Computer

By all accounts, the rst machine to include rhythm quantization as an out-of-the-box, usable feature was C17.P12
8
Roger Linn’s LM-1 Drum Computer (see Figure 17.2). Released in 1980 with a $5,000 price tag, the novelty
p. 346 of this device hinged on two key technical innovations. First, whereas other products on the market at
that time merely simulated drum sounds using analog synthesis techniques, the LM-1 used 8-bit digital
9
samples of ‘real drums’ (as the advertisement prominently asserts). And second, unlike earlier drum
machines, Linn’s device allowed users to program rhythms in ‘real time’, as well as to ‘auto-correct’ and/or
‘shu e’ rhythms by moving them in relation to a metric grid. If the use of digital samples was remarkable,
it was the second set of features that transformed rhythm in popular music. The device’s ‘auto-correct’
function o ered options for setting the quantum value to eighth, sixteenth, or thirty-second note
p. 347 durations, in either duple or triple subdivisions; in addition, its ‘shu e’ (or ‘swing’) function o ered six
settings that ranged from 50% (straight) to 70% (heavily swung). Playing with just these variables, users
exercised their expressive control over the quantization process, producing tightly synchronized beats that
could be reanimated with a ‘human rhythm feel made possible by special timing circuitry’.
Figure 17.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Advertisement for Roger Linnʼs LM-1 Drum Machine (c.1981). To view the figure in colour, please visit the companion website C17.F2

Since its release nearly four decades ago, the LM-1 Drum Computer has garnered an almost mythical status, C17.P13
especially in the world of music techies and nostalgic audiophiles. This status can, at least in part, be
attributed to the LM-1’s novel functionality and its early promotion by several key musicians, including
Stevie Wonder, Prince, and Michael Jackson (McNamee 2009). But the public perception of the LM-1 as a
cult object has also been nurtured by a steady stream of press hyping its serendipitous origins. According to
Linn’s own version of events as recounted in several interviews, the ‘auto-correct’ and ‘shu e’ features
were more of a discovery than an invention. Here’s Linn describing the genesis of rhythm quantization:

Some of my best innovations were discovered by accident. I created the quantize feature while C17.P14
trying to nd a software algorithm for real-time sequence recording that conserved expensive
computer memory […] when I initially tried recording a 16th-note drumbeat, I noticed that the
software was moving my notes onto the nearest 16th-notes. The software was correcting my
timing errors!

(Linn 1998)
In this account, quantization was not the result of an intentional goal, but rather a consequence that C17.P15
emerged from the technical constraints of the device’s infrastructure. An unforeseen lack of operating
power restricted the machine’s capacity to process real-time rhythmic input, and then, through a stroke of
luck, Linn discovered quantization, forever altering the course of popular music. Admittedly, this is an
attractive narrative that imbues the LM-1 Drum Computer with a special element of providence; but it
seems unlikely that such a momentous development could have happened so haphazardly and in such an
isolated context.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Since rhythm quantization is purported to have been invented by a single person working alone, it is not C17.P16
easy to con rm that Linn’s LM-1 is indeed the origin of the technology. This situation is made even more
di cult by the fact that, in a puzzling omission, Linn’s 1979 patent application for a ‘Modular Drum
10
Generator’ lacks any claims about the device’s ‘auto-correct’ or ‘shu e’ functions. To gain some clarity, I
contacted Linn inquiring about the matter: what exactly were the circumstances surrounding the accidental
discovery of quantization? And why was it not included in his patent application? In his response, Linn
explained that it was di cult to acquire patents for software at that time, and as a 22-year-old with limited
11
funds, he had ‘bigger sh to fry’. And as for quantization, Linn conceded that the widely circulated story
about its discovery was largely intended for anecdotal e ect, ‘In truth, I knew that correction of timing
errors was needed but the accidental discovery makes for a better story […] That said, when I rst tried it out
12
I was surprised by how well it worked.’

p. 348 In hindsight, it stands to reason that Linn would have already known about the need to quantize rhythmic C17.P17
input, since the makers of Roland’s CompuRhythm CR-78 had already realized the utility of the technique
two years earlier out of sheer necessity.

The CompuRhythm CR-78 (1978) was an analog drum machine that enjoyed widespread popularity C17.P18
amongst acts such as Peter Gabriel, Genesis, and Ultravox. Like the LM-1, the device could be programmed
using a real-time input method, but its lack of a quantization feature made it di cult to produce tightly
synchronized drum patterns. To remedy the situation, Roland subsequently released the WS-1 write switch,
13
an accessory that made it possible to divide ‘the rhythm pattern into steps of either 8-12-24 or 48’. This
may seem like a minor detail, but the ad hoc assemblage of drum sounds, real-time input methods, and a
14
scalable metric grid found in Roland’s CR-78 and WS-1 sheds new light on the origins of quantization.
Looking back, it appears that the implementation of quantization in the LM-1 was more of a re nement and
repackaging of previous technological processes, rather than a sui generis invention. And the picture grows
even more complicated if one considers the isolated origins of quantization’s individual components. For
instance, real-time input methods had already been introduced as part of the build-your-own PAiA
Programmable Drum Set (1975), and grid-based programming had premiered in machines like Eko’s
ComputerRhythm (1972), which featured a six-row push-button matrix that pre gured later step-
15
sequencers like the Roland TR-808 (1980). These machines acted as vital stepping-stones, without which
Linn’s LM-1 Drum Computer could not have been built. Their role as antecedents to the LM-1 undercuts
prevailing narratives about the singular origins of rhythm quantization, suggesting instead that it emerged
gradually through a coalescence of analog and digital sequencers and synthesizers, linked through a
piecemeal series of technological obsolescence and displacement.
Interlude on the Paradoxical Nature of Rhythm Quantization C17.S3

Before leaving the LM-1 behind, it is worth pausing to consider one more curious detail about the device’s C17.P19
interface. As shown in Figure 17.2, the machine’s auto-correction module included a ‘HI’ setting in addition
to the other quantum values. This high-de nition mode for programming disabled the machine’s auto-
correct function, preserving what the advertisement billed as the ‘real time’ of the original performance.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Technically, the timing was still constrained by the pulse rate of the device’s sequencer, which was limited
to 48 parts per crochet. And while this was fast enough to evade perception, it was hardly what we’d
consider ‘real time’ today. In any case, this ambiguity points to a separate but related meaning that
quantization has in the context of digital audio, where it refers to an inherent reduction of information that
16
occurs when analog signals are digitized. At base, analog-to-digital conversion involves three essential
p. 349
steps: (1) sampling a continuous audio signal, (2) quantizing the amplitude of each sample into a discrete
value, and (3) coding the quantized data into binary notation. In the second phase of this process, the
discrete amplitude resolution of each sample will inevitably di er from the original waveform, introducing
a certain amount of what is known as quantization error or noise into the digital signal. Developers of
telecommunications technologies have generally sought to reduce this interference as much as possible,
striving for an impossible state of unmediated transparency and holding to the ideal of delity to an original
17
signal. But as the language of digital audio quantization has ltered into discourse on rhythm
quantization, the connotative meaning of the word has changed, with verisimilitude discarded and the
introduction of noise into the signal embraced as a desirable quality. In this sense, when quantized rhythms
are perceived as being perfectly timed, it is because the listener is aestheticizing the space between the
original performance and its quantized representation.

The semantic reorientation outlined here highlights the often contradictory values that operate within what C17.P20
software theorist Lev Manovich describes as the dual layers of new media objects: a cultural layer and a
computer layer (2001: 44). In the former, the representation of rhythm remains recognizable at the
vernacular level of human culture, whereas in the latter, rhythms are rendered as machine-readable code,
18
entering into dialogue with the computer’s language of algorithms and data structures. The interaction of
these two layers gives rise to an apparent contradiction, in which something that is regarded as an ‘error’ at
the level of computer technology can be simultaneously perceived as a ‘correction’ at the level of human
culture. In the case of musical time, this perceptual incongruence occurs when a culture’s musical bearings
are redirected away from subjective conceptions of temporality and towards the objectivity of a
chronometric reference. Because such a shift was necessary for the ascendancy of rhythm quantization, we
can be sure that the cultural antecedents of quantizing processes extend further into the past than its
technological ones.
Cultural Preparation for a Measured Approach to Musical Time C17.S4

Long before Linn’s LM-1 Drum Computer, the technocultural foundations were laid for rhythm C17.P21
quantization to become a viable musical practice. The most obvious instruments of change during this
process of social regimentation were click tracks in the context of lm music and multi-track recording
studios, and metronomes in performance practice more generally. Using these simple tools, performers

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
began to measure musical time against the backdrop of a mechanical beat that corralled players into the
p. 350 con nes of metric conformity. In the case of click tracks, their initial use in lm during the 1920s–1930s
derived from practical necessity: producers were searching for a way to synchronize musical tempo with the
pace of action on the screen. As Thomas Cohen recalls, in the early days of click tracks, producers would
punch holes at regular intervals along the side of a piece of lmstrip, causing it to make a clicking sound
during playback (Cohen 2009). Fast-forwarding a few decades, the click track became an invaluable tool in
multi-track studios, where it was used to synchronize audio channels recorded independently of one
another. According to Paul Théberge, the rise of multi-track studios ushered in a new epoch in which the
‘process of group performance and the social/musical exchange between musicians became rationalized
and fragmented—both spatially and temporally’ (1997: 216). This fragmentation process transformed the
traditional notion of music-making as a social activity grounded in the physical co-presence of the
participants. In its place, it presented listeners with a simulated ensemble cobbled together from individual
performances. More recently, the advent of the World Wide Web has further ampli ed this aspect of
fragmentation thanks to the rise of ‘network studios’, which facilitate the global redistribution of music
production across a number of remote studio outposts communicating with each other within a ‘virtual
space’ (Théberge 2004). With performers spread over such wide expanses, it has become a matter of
necessity for producers to institute a global reference that maintains temporal order within a production
process that has become increasingly discombobulated.

Click tracks themselves were not without precedent; they merely extended the functionality of metronomes C17.P22
beyond the practice room and into the studio or concert hall. In this respect, the widespread production and
distribution of Johann Maelzel’s ‘metronome or musical time keeper’ (1815) must be counted as an
important historical moment because it helped to reorient popular conceptions of musical time towards a
newfound objectivity ensured by mechanical agents. Unlike earlier chronometers, the clockwork
metronome inaugurated a universal standard for the measurement of tempo in beats-per-minute,
signalling a shift away from subjective timing as performers disciplined their actions according to the
authority of an external referent. In Chapter 16 of this Handbook, Alexander Bonus details how the
metronome was, and still is, greeted with scepticism by many who have viewed it as an a ront to the
musician’s interpretive license for expressive timing. Even Maelzel himself warned that one should never
‘play an entire piece with the metronome because all expression would be paralysed by such slavery’
(Maelzel 1822: 5). But despite this word of caution, metronomic timing long ago assimilated into
performance practice, setting standards for acceptable tempos that are enforced by conductors, spread by
recordings, and internalized by working musicians. The history of metronomes thus provides necessary
perspective on a fundamental tension that lies at the heart of rhythm quantization: that between regulative
and a ective expressions of musical time. Echoing contemporaneous arguments around mechanism and
vitalism in the physical sciences, the dichotomy between regulative and a ective timing in music sparked
erce debates throughout the nineteenth century, as the proliferation of the metronome began to impinge
p. 351 on perceived relationships between individualized expression, physical embodiment, and cultural
authenticity.

According to Myles Jackson, the gradual formation of a consensus around beats-per-minute as a universal C17.P23
measurement for tempo involved multilateral negotiations between acousticians, physicists, and mechanics
from the science domain, and performers, composers, conductors, and publishing houses from the musical
domain (Jackson 2006). In addition, the consensus took shape in parallel with a gradual standardization of
tuning, which led to the adoption of an international European standard in which a4 was equal to 435 Hertz
(2006: 224). Scaling upwards, these dual trajectories towards standardization of musical time and pitch can
19
be viewed against the backdrop of a broader epistemological shift towards mechanical objectivity. This
shift was fuelled by a growing distrust of subjectivism and, as new tools and standards of measurement
became available, an increasing reliance on numbers and scienti c evidence that sought to represent the
world through veri able facts. Within the context of music, the push for greater objectivity collided with

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
existing sensibilities for musical style, and as a result, debates around tempo and tuning often centred on
the relationship between standards and aesthetics. As debates wore on, these two aspects became
increasingly inseparable, such that the ability to measure time perfectly began to shape public perception of
what counts for good musical timing.

The embrace of mechanical objectivity in musical practices was signalled not only by the mechanical C17.P24
metronome, but also by an in ux of musical automata dating back to the eighteenth century. Interestingly,
metronomes and automata both passed through the workshop of Johannes Maelzel, and it’s worth
considering how these two sides of his enterprise relate to quantization. As Jessica Riskin (2003) argues, the
wave of mimetic automatons unleashed during the eighteenth and nineteenth centuries aimed to
understand life by simulating it in mechanical form. Over time, an ongoing dialectic emerged between
machines and bodies, in which the limits of mechanization were repeatedly dramatized by assigning tasks
to automata that were thought to be distinctly biological in nature. There were speaking machines,
performing musicians, and even defecating ducks, all of which raised the ontological question of semblance:
for if machines could simulate biological processes, then what exactly distinguished animals from
machines? As the boundary between humans and machines became increasingly blurred, the mimetic
potential of automata triggered a reappraisal of the human condition and a subsequent reversal of the man–
machine metaphor such that, instead of machines emulating human behaviour, people began modelling
20
themselves on machines. While it is impossible to pinpoint the precise moment of this referential twist in
the history of musical time, it was most certainly prior to Edgard Varèse’s musings in 1939 about a ‘sound-
21
producing machine’ capable of producing rhythms that are ‘humanely impossible to attain’ (1966: 14). In
more recent years, the persistence of this trend can be witnessed in the rise of technomorphism, with
22
composers turning to the electronic music studio as a source of new models for instrumental composition.

If one digs deep enough into the history of rhythm quantization, there comes a point at which material C17.P25
p. 352 objects and artefacts dissipate, leaving the grid-like structure of the metric hierarchy intact and free-
standing without the aid of technical support. It is within this hierarchical conception of musical metre that
one nds the deep roots of quantization intertwining with larger debates about the epistemological
grounding of time itself. As Roger Mathew Grant argues (2014), a major shift in theories of rhythm and
metre during the eighteenth century precipitated the emergence of a hierarchical perspective on musical
time, and one of the earliest and fullest expressions of this new perspective can be found in Johann Philipp
Kirnberger’s 1779 treatise. Prior to this point, the mensural system had contextualized the beat, or tactus, in
relation to the down–up, thesis–arsis motion of the hand, or alternately, through the twofold motion of the
beating heart (Berger 2002). But for Kirnberger, the assumed correlation between time and embodied
motion gradually gave way to a more objective conception of musical time, which he described as follows:

If one hears a series of equal beats that are repeated one after the next with an equal space of time C17.P26
[…] experience teaches us that we, in our minds, immediately divide them metrically by arranging
23
them in segments, each containing an equal number of beats.

One of Kirnberger’s key contributions lies in his description of a pre-existing series of undi erentiated C17.P27
pulses detached from all musical context. This view broke from previous theories, and, as Carl Dahlhaus has
noted, it closely mirrored Isaac Newton’s earlier formulation of absolute time as that which ‘ ows equably
24
without relation to anything external’. As such, Kirnberger’s theory was central in establishing a more
objective basis for musical time, harnessing Newtonian pulse as a stable foundation for the construction of a
metric hierarchy that remains more or less intact today.

In light of this long succession of temporal regimes, we can understand the relatively recent rise of rhythm C17.P28
quantization as just one more step in a long line of regulators, following behind mechanical clocks,
25
pendulums, hour glasses, and various other means for measuring time. Indeed, a reciprocal loop can be

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
drawn between these objects, theories of time as a scienti c object of study during the Enlightenment, and
the aesthetics of musical time as developed shortly thereafter by Kirnberger and others. Flashing forward,
the rings of this loop widen, such that the inner workings of a device like the LM-1 Drum Computer forge a
link between musicians and time that is conditioned by an underlying system of metric hierarchies. The end
result is that, when quantization algorithms determine which way to nudge a rhythmic unit by assuming its
position within a regular succession of undi erentiated pulses, they are silently rehearsing the history of
debates around mechanization and abstract notions of absolute time. Thus, despite the accusations of
dehumanization that have accompanied rhythm quantization, the process is in fact grounded in beliefs
about time that are entirely human because, without the cultural preparation of a metric hierarchy and
Newtonian pulse, there would be no basis for the ensuing aestheticization of quantization in popular music.

p. 353
Genre-Based Reception History of Rhythm Quantization C17.S5

Over the past 30 years, rhythm quantization has become a pervasive practice in popular music production, C17.P29
eliciting both anxiety and adulation depending on the genre. Initially the province of disco, electronica, and
other genres that embraced technology as part of a particular aesthetic orientation, the in uence of
quantization has since spread across a broad spectrum of musical styles. Today, even musicians who may
have once opposed quantization will acquiesce to small doses, catering to market demands for polish and
precision on studio albums. As a result, the dividing line between those who do and don’t use quantization is
no longer clear-cut. The technique can be applied in many ways—either explicitly for its own mechanizing
e ects or as a barely noticeable touch-up—and the decision of whether and how to use quantization
depends entirely upon a given genre’s shifting codes and customs. In what follows, a paratextual analysis of
statements by some of the main actors within the music industry provides evidence of wide variance in the
public reception of quantization, suggesting that there is still much at stake in ongoing debates around this
evolving practice.

Discourse on rhythm quantization frequently centres on the aforementioned tension between a ective and C17.P30
regulative expressions of musical time, and this dichotomy tends to weigh heavily on whether or not a
performance is perceived as being ‘authentic’. As Simon Frith notes, many people view technology as
intrinsically opposed to authenticity because of a belief that it either obstructs communication between the
26
artist and audience, or else implies cheating on some level (Frith 1986). A prime example of this viewpoint
can be found in James Brown’s critique of disco as a genre that was ‘easy for artists […because] it was all
electronic sequencers and beats-per-minute—it was done with machines. They just cheated on the music
world’ (Brown and Tucker 1986). While this criticism is perhaps to be expected from the ‘hardest working
man in show business’, Brown is hardly alone: his critique resonates in many corners of the music world,
tapping into broader anxieties about the impending loss of ‘liveness’ in performance and the spectre of
27
human expression being supplanted by machines. From this perspective, quantization is regarded as an
invasive force that strips music of its unique aura, resulting in the erasure of what Charles Keil calls
‘participatory discrepancies’—those minor tears and ripples in the seams of time that result from the inter-
subjective nature of group performance (Keil 1987). Concerns about the loss of artistic aura are nothing new,
of course, but the heightened e ciency with which quantization imposes chronometric uniformity has
given fresh force to long-running arguments against the standardizing e ects of mass industry. One can
only imagine what Adorno might say, given his polemics against popular music and his disparagement of
‘those individuals of the rhythmically obedient type […who] are most susceptible to a process of masochistic
28
p. 354 adjustment to authoritarian collectivism’. This old line of attack persists today in slightly varied form,
emerging as a perennial defence against what is perceived to be the homogenizing rhythmic in uence of
quantization on mainstream musical culture.

Fears about the incursion of machines into music performance derive as much from the lived social and C17.P31

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
economic realities of working musicians as they do from any general aesthetic considerations. In this
regard, Brown’s critique re ects widespread concerns about the displacement of human labour by an in ux
of new technologies during the 1970s. As Sarah Angliss (2013) has shown, anxiety over automation
consumed organizations like the United Kingdom’s Musicians’ Union, which launched a full-scale
campaign in 1982 against what it considered to be ‘canned music’. During the course of this campaign, a
blanket ban was placed on all drum machines (including the LM-1), samplers, and synthesizers, which were
not allowed into concerts and recording sessions for fear of putting musicians out of work. While it’s true
that, in some cases, these machines had a detrimental impact on existing practices and social
con gurations within the music industry, they also enabled the formation of new subgenres (and entire
subcultures) that responded to the realities of a post-industrial landscape. The inability of the Musicians’
Union to foresee the rising popularity of these technologically driven subgenres all but guaranteed the
defeat of their e orts to ban electronic music devices from the industry.

From the other side of the debate, fans of disco and early electronic music tended to embrace technological C17.P32
arti ce as a rebuke to the perceived excesses of rock ’n ’roll. Writing for the Village Voice at the height of the
disco era, Andrew Kopkind re ected on how ‘disco in the ’70s is in revolt against rock in the ’60s. It is the
antithesis of the natural look, the real feelings, the seriousness […] disco is “unreal,” arti cial, and
exaggerated’ (1979). In analogous fashion, electronic groups such as Kraftwerk traded their guitars for
29
drum machines and synthesizers as part of a conscious e ort to cultivate a machine aesthetic. More than
mere shock art, their preference for mechanized expressions opened out onto a larger worldview, in which
creative musicians were recontextualized as ‘musical workers’, and the e orts of both humans and
machines were dissolved in a tableau of techno-futurist themes and science ction motifs, all of which
heralded the ltration of post-humanist thought into the substrate of popular culture.

Adopting a similar position, early techno DJs in Detroit, such as Cybotron and Drexciya, harnessed the C17.P33
characteristic rigidity of quantized rhythms as a sonic corollary to their construction of a distinctly Afro-
futurist cosmology. Within this social and political context, quantization engendered new modes of
expression and helped to establish sites of resistance to normative notions of authenticity. British music
critic Kodwu Eshun captured the essence of this zeitgeist when he proclaimed that ‘the posthuman era is not
one of disembodiment but the exact reverse: it’s a hyper-embodiment’ (1998: 2). Rejecting the primacy of
corporeality in naturalized expressions of soulfulness, this assertion advances the cause of a Black post-
humanism, which Eshun o ers as a riposte to overly restrictive constructions of human subjectivity in
Western philosophical discourse. His argument runs parallel, in this regard, to one made by Alexander
Weheliye, who focuses on the use of vocoders in contemporary RandB to show how digital technologies have
p. 355 allowed musicians to ‘reconstruct the black voice in relation to information technologies […creating] a
composite identity, a machine suspended between performer and producer that sounds the smooth ow
30
between humans and machines’ (2002: 30). The subversive foregrounding of mediated vocal expressions
in post-soul music has its counterpart in rhythm quantization because, like the voice, discourse on the
relation of human ‘feeling’ to rhythm often veers into the realm of essentialist stereotypes—a fact borne
out by the long history of popular music criticism that conceptualizes swing and groove in blatantly
racialized terms.

These polarizing views around quantization in its early years make clear that what counts as authenticity in C17.P34
musical expression is a highly contentious issue, and one that can be oriented around vastly di erent value
systems. But as quantization became an increasingly common practice in digital recording studios, many of
these tensions subsided and opposition to its use gradually diminished, causing its in uence to spread into
the mainstream. By the 1990s, there appears to have been a general acceptance in many genres that some
amount of quantization was necessary as a corrective on studio albums. This was true even in the case of
rock, where overt displays of physicality had long been central to the identity of the genre. Re ecting on this
trend, producer Bob Rock recently exclaimed:

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Everybody’s just got to lighten up on the grids and the Auto-Tune. That’s why the White Stripes or C17.P35
the Strokes or even St. Anger [by Metallica] are so abrasive to people. Because they’re not actually
that abrasive. Those kinds of records have been around forever. It’s just that now, when everything
is perfectly in time and perfectly in tune, that’s what people expect.

(quoted in Battino and Richards 2005: 137–138)

This lament for a bygone era when recordings captured live performances in a supposed state of raw C17.P36
expression seems to be emblematic of general attitudes towards technology in rock music no matter how
31
erroneous the claim that recordings ever existed in a ‘raw’ or unmediated state. From Bob Dylan’s famous
‘Judas!’ moment (1966) to the critical blowback Radiohead received after embracing electronica on albums
like Kid A (2000), rock audiences have registered persistent unease with the introduction of new sound
technologies into established musical contexts. Couched within this historical discontent, Bob Rock’s
statement re ects a common frustration with the way digitization has altered the sounds of contemporary
rock, and it speaks to a growing fear that quantization has initiated an irreversible loop or cultural shift,
with the sensibilities of younger fans recalibrated towards an exaggerated studio perfectionism that can
only be maintained through extreme technological intervention.

Against the relative orthodoxy of rock ’n ’roll, there has been a more enthusiastic acceptance of C17.P37
quantization among many heavy metal musicians, especially in subgenres such as ‘Djent’ (so-named for its
characteristic use of percussive power-chords). To wit: in a recent manual on how to produce a heavy metal
album, Mark Mynett discusses the proper use of what he calls ‘gridding techniques’, advising that, ‘when
producing bands in this style of music, the most suitable approach to provide the pre-requisite ethics of
p. 356 precision, accuracy and tightness should be given priority over the reality of the drum performance
event’ (2013: 201). In other words, given the genre’s emphasis on rhythmic precision, producers should err
on the side of more quantization, trading delity for an exaggerated virtuosity that verges on hyper-
realism. This position contrasts sharply with traditional rock music: instead of reluctant acceptance, certain
subgenres of extreme metal are so heavily invested in quantization that it has become a constitutive
precondition of the music itself.

Another stylistic context in which quantization has played an important role is early hip hop and rap music. C17.P38
Throughout the 1980s, producers readily adopted the use of digital drum machines and samplers, and
because quantization was a stock feature on popular devices like Akai’s MPC60 (1988), it was practically
hardwired into hip hop recordings from this period, achieving the same kind of ubiquitous presence it
32
enjoyed in other genres. This in uence began to wane in the late 1990s, however, as experimental hip hop
producers increasingly shirked the conventional wisdom that mandated beats be properly quantized.
Instead, artists such as J Dilla began pioneering approaches to beat-making that de ed easy categorization.
The characteristically ‘wobbly, seasick’ grooves that resulted from these alternative strategies have been
33
hugely in uential, not only for younger producers but for live drummers as well. As an example of the
latter, Questlove (drummer for The Roots) recently recounted the story of how, after initially training
himself to play like he was ‘just an absolute machine’, he ‘met D’Angelo and J Dilla, who then made me just
dismantle everything that I knew about being cold’ (2013). Here, we have a performer who is trying to mimic
a drum machine, which itself is manipulated to mimic the microrhythmic instability of a (possibly tipsy)
human drummer. As this twisted chain of imitation makes clear, rhythm quantization has been thoroughly
massaged into all layers of musical culture, extending from the digital domain back into the realm of
acoustic instruments and live performance.

Despite the prevalence of quantization in popular music, some artists continue to resolutely reject the C17.P39
technology as anathema to their aesthetic values. Take the case of jazz: the importance of swing in jazz is a
prime example of how microrhythmic expressivity can be a de ning factor in a genre’s identity. Several
researchers in the burgeoning eld of performance studies have endeavoured to investigate the irreducible

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
aspects of rhythm in jazz (Benadon 2006; Butter eld 2006; Do man 2013), employing empirical methods
to examine the way performers play against the metric grid, exing its frame. Likewise, similar approaches
can be found in studies of classical music, such as Bruno Repp’s comparison of 19 performances of a minuet
by Beethoven (1990) or, more recently, Mitchell Ohriner’s examination of listener–performer synchronicity
in recordings of Chopin’s Mazurkas (2014). Interestingly, the results of these studies are not just relevant
for those wishing to demonstrate the importance of expressive timing; they have also been used by
computer scientists to calibrate quantizing operations in performance software applications. A notable
breakthrough in this regard can be traced to Edward Large (2000), a psychologist whose work inspired
researchers at the Parisian Institut de recherche et coordination acoustique/musique to develop a score-
following software that synchronizes live electronics and performers in the context of mixed music (Cont
p. 357 2008). In this case, quantization is actually being harnessed for purposes that run counter to its original
intent. Rather than adjusting the performer’s rhythm to the computer’s internal clock, it does just the
opposite: deciphering a performer’s expressive deviations from an underlying metrical structure, this type
of software e ectively warps the computer’s grid to the performer, subtly adjusting the tempo in the same
way that a good accompanist might. As a result, score-following algorithms bolster the live performer’s
ability to take expressive liberties, enacting a reversal of rhythm quantization that is tailored towards
decidedly di erent aesthetic aims.

Looking Ahead at the Cross-cultural Politics of Rhythm Quantization C17.S6

Thinking back to Glen Ballard’s comment about the ‘authority’ of rhythm quantization, it is essential to C17.P40
consider how this subliminal power operates in di erent cultural contexts, spreading via the proliferation
of commercially produced and globally distributed music technologies. This is still a neglected area of
research, but a growing number of authors have begun to take notice. Among them is Martin Scherzinger,
whose work details the incongruences that arise between Western musical values embedded within groove-
tracking software and those of non-Western cultures. Fusing aspects of musicology with media studies,
Scherzinger has proposed a software physiognomics that ‘maps the way software remodels the a ective life
of individuals’ (2016: 68). He has also suggested that the encoding of perceptual analytics within audio
engineering standards has led to a situation where, in e ect, ‘the machine does the listening for you’ (p.
34
70). Extended to quantization software, which is productive (and not merely re-productive), this implies
that the machine also does the performing for you, not just listening. As such, it becomes all the more
critical to excavate the cultural biases embedded within software algorithms, calling attention to their
contingency in order to develop alternative programs that might encourage more inclusive coding
35
practices.

In many ways, studies on the hidden cultural dimensions of quantization software have their antecedents in C17.P41
historical debates around the politics of transcription in ethnomusicology. This relationship stems from the
fact that, like software, notation acts as an analytic mediator, reducing sounds to symbols that signify
whatever musical values were deemed important by the system’s developer. Writing in 1958, Charles Seeger
was already keenly aware of this problem, arguing that the prescriptive nature of western notation, which is
primarily concerned with discrete structures of pitch and rhythm, was wholly inadequate for the purposes
of descriptive transcription. Consequently, ltering non-Western music through the sieve of Western
notation not only produced a misleading representation of the analytic object but also imposed a ‘riot of
36
subjectivity’ onto the culture being studied (Seeger 1958: 187). Perhaps something similar can be said
about the way quantization imposes Western values on popular music; even more so because of its
p. 358 concealed nature and its audible e ects that transcend mere symbolism. For this reason, debates on the
ethics of transcription can be read in parallel with the present study, shedding further light on the cross-
cultural politics of rhythm quantization.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Clearly, there is much work to be done if we are to arrive at a better understanding of the countless C17.P42
contingencies upon which quantization is founded. This chapter has only scratched the surface, attempting
to map the broad contours of a new area of study by showing how quantization operates at di erent
registers of musical culture. To this end, I have sketched a sprawling network of objects, ideas, people, and
institutions, presenting a number of views both for and against quantization. In addition, I have included
accounts from musicians who reside in the spaces between these extremes, either using quantization in a
sparing manner or repurposing it towards di erent ends. The sheer diversity of musical applications in
existence attests to the fact that quantization is a malleable technology that can be moulded to t vastly
di erent artistic intents and purposes. Given this relativity, there is no sense in arguing over whether or not
quantization can be reconciled with this or that version of musical authenticity; and in fact, the terms of this
debate can be rejected altogether on the grounds that such binary choices only help to reinscribe untenable
divisions between technology and nature. As I have argued, quantization has cultural origins that are
entirely human, and as such, it should be understood as an externalization of social and cognitive processes,
not an alien imposition into human a airs. Extending back to antiquity, quantization has been preceded by
a lengthy succession of time-keeping techne, which were themselves encapsulations of earlier technical
components, abstract conceptual schemas, and the wider cultural milieu within which they were incubated.
Hence, whenever quantization is applied in a musical situation, it silently invokes this long history, re-
enacting the entire series of scienti c and philosophical debates outlined earlier in this chapter. At the same
time, however, it adds something new by transcoding these established cultural forms into a media system
that engenders its own distinctive set of sonic practices. Operating on symbols instead of sound waves,
digital quantization allows users to cleave apart musical surfaces and radically recon gure temporal
relationships. In doing so, it normalizes and rei es the metric grid, magnifying its in uence by making it
the universal backdrop against which all musical time is measured. Whatever the aesthetic rami cations of
this universalizing impulse, the fact remains that quantization has become and will remain a constitutive
force within popular music culture—one that must ultimately be regarded not as an authority governing
unmediated musical expression, but rather as an extension of social and material realities that were forged
in historical memory long before they were encoded in digital sound technologies.

Notes
1. Ballard o ered this assessment in a panel discussion on rhythm quantization with several notable figures from the music C17.N1
p. 359 industry, including Roger Linn (drum machine maker), Nile Rodgers (producer), Gerhard Behles, and Robert Henke (co-
creators of Ableton Live so ware). See Battino and Richards (2005: 134).

2. For a technical description of rhythm quantization, as well as a survey of various mathematical and computational C17.N2
approaches to the problem, see Boenn (2018: 147–173); and for a practical explanation of how quantization works in the
context of digital studio production, see chapters 3 and 4 in Pejrolo (2011).

3. MIDI (ʻMusical Instrument Digital Interfaceʼ) refers to a standard protocol for connecting electronic music devices. In Figs C17.N3
17.1b and 17.1c, a device known as a MIDI controller was used as an input method for programming the rhythm in Ableton
Live. For more on the history of MIDI, see Diduck (2017).

4. The interface shown here is specific to Ableton Live, but most other digital audio editing so ware, such as Logic Pro or Pro C17.N4
Tools, adhere to similar design principles.
5. For a more detailed description of these quantization factors, see the Ableton Reference Manual Version 9 (DeSantis et al. C17.N5
2016: 137–182). And to learn more about recent attempts to ʻhumanizeʼ digitally programmed rhythms, see Holden (2015)
on his development of a ʻGroup Humanizerʼ plug-in for Ableton Live. This virtual device goes beyond randomization,
aiming to simulate the microrhythmic entrainment found in a performing ensemble via cognitive models outlined in a
recent Harvard study on ʻSynchronization in human musical rhythmsʼ (Hennig 2014).

6. See Agon et al. (1994), Cemgil et al. (2000), and Desain and Honing (1992). C17.N6

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
7. See e.g. Lerdahl and Jackendo (1983), Todd (1985), or Clarke (1987), all of which assume a hierarchical separation of C17.N7
structural versus expressive musical time in classical music, and have been credited in subsequent so ware development
by Desain and Honing (1989), Agon (1994), and others.

8. See e.g. (2016), Mask (2014), or Scarth and Curry (2013). According to these accounts, rhythm quantization began with the C17.N8
release of Linnʼs LM-1 and then quickly became an industry standard, as evidenced by the proliferation of competing drum
machines and samplers with quantizing functionality, including Oberheimʼs DMX (1981), Yamahaʼs RX-11 (1984), and Akaiʼs
MPC60 (1988).

9. According to Linnʼs ʻPast Products Museumʼ, the real drum sounds were ʻmostly recorded by local L.A. session drummer Art C17.N9
Wood, and sampled at only 28 kHz using an 8-bit non-linear formatʼ (Roger Linn 2018). The pictured advertisement dates
from April 1981, and was printed in numerous music technology magazines, including Keyboard.

10. See Linn (1979). The primary objects of Linnʼs invention were limited to the deviceʼs use of digital samples and its ability to C17.N10
be remotely controlled and linked to other machines.

11. Roger Linn, correspondence conducted via email, 26 Sept. 2017. A quick internet search confirmed Linnʼs assertion, as the C17.N11
U.S. Patent and Trademark O ice did not start routinely issuing so ware patents until sometime a er the 1981 Supreme
Court case, Diamond v. Diehr (https://www.bitlaw.com/patent/index.html).

12. Ibid. C17.N12

13. For more on the WS-1, see the Roland CR-78 ownerʼs manual (1980). While a precise date for the release of the WS-1 is C17.N13
unavailable, it was manufactured during the same year as the CR-78 (1978) and released a few months a erwards. For
more information, see Vail (1994).

14. The argument for a distribution of agency is further supported by Rolandʼs founder, Ikutaro Kakehashi, who said the C17.N14
following about the creation of Rolandʼs CR-78: ʻMany people were engaged. The Japanese system takes team work, so
p. 360 there wasnʼt one person who was responsible. This is quite di erent from the US system. In the US, very clearly, “I
designed everything” is what many people say. But in Japan, no. So itʼs very di icult to say who did whatʼ (quoted in Vail
1994).

15. Interestingly, Ekoʼs ComputerRhythm also read punch-cards, linking the machine back to earlier forms of material C17.N15
automation such as those found in player pianos. For more information on these earlier drum machines, see Wilson
(2016).

16. Quantization in this context can be traced back to telephony research conducted at Bell Labs during the 1930s–40s and C17.N16
summarized in a seminal 1948 article on pulse-code modulation, ʻThe philosophy of PCMʼ, co-written by Oliver, Pierce, and
Shannon.

17. Impossible because, as Jonathan Sterne argues, the concept of sound fidelity has always been ʻmore about enacting, C17.N17
solidifying, and erasing the relations of sound reproduction than about reflecting on any particular characteristics of a
reproduced soundʼ (2003: 274).

18. In actuality, both are ʻhumanʼ layers, but one runs on the semantics of machine operations, whereas the other is C17.N18
understood in relation to the cultural form it is meant to represent.

19. Writing about the evolution of conventions around the visual representation of objects in scientific atlases, Daston and C17.N19
Galison have shown how there was a gradual shi from representations based on an ideal of ʻtruth-to-natureʼ in the 17th
and 18th cc. to a new paradigm that aspired to ʻmechanical objectivityʼ in the 189h c. For more on the shi to mechanical
objectivity, see chs 2 and 3 in Daston and Galison (2010).
20. Thinking of human behaviour in technological terms reached its highpoint in the mid-century development of cybernetics C17.N20
by Norbert Weiner and others, who posited that ʻthe ultra-rapid computing machine […] must represent almost an ideal
model of the problems arising in the nervous systemʼ (1948: 14).

21. The irony of this otherwise prescient comment is that such a machine had already taken shape; in 1931, Leon Theremin C17.N21
built the ʻRhythmiconʼ, foreshadowing what would become a steady stream of drum machines in the coming decades
(Schedel 2002).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
22. Technomorphic tendencies are especially pronounced in certain styles of contemporary music, such as French C17.N22
spectralism (Wilson 2000), and one finds analogues for this shi in numerous other fields, including architecture (Nero
2004) and psychology (Waters 1948).

23. Kirnberger (1982[17769]: 383). C17.N23

24. Newton (1999[1687]: 62). And for more on the relationship between Newtonʼs theory of absolute time and Kirnbergerʼs C17.N24
development of Akzenttheorie, see Dahlhaus (1984: 160).

25. According to Mumford, the impulse to regulate time can be traced back to medieval monasteries, which instituted the C17.N25
practice of ringing bells seven times a day to mark the canonical hours. In the absence of clockwork, these bells lent
ʻhuman enterprise the regular collective beat and rhythm of the machine; for the clock is not merely a means of keeping
track of the hours, but of synchronizing the actions of menʼ (1934: 14).

26. Following Frith (1986) and others, I do not regard musical ʻauthenticityʼ as something that resides in the music itself; C17.N26
rather, it is a contingent notion that is mutually constructed by artists and audiences working within the context of a
specific historical and cultural milieu.

27. According to Auslander (1998), ʻlivenessʼ is an ideological construct that emerged during the 1930s in response to a crises C17.N27
sparked by the rise of radio broadcasting, in which in-studio performances were indistinguishable from recordings. In this
sense, defenders of ʻlivenessʼ may be surprised to learn that their favourite rock albums su er from an incurable
ʻdeadnessʼ (for more on the concept of ʻdeadnessʼ in sound reproduction, see Stanyek and Piekut 2010).

p. 361 28. Adorno makes this hyperbolic assertion in relation to repetitive dance rhythms in so-called ʻjitterbugʼ music of the 1930s C17.N28
and 1940s, which he deemed symptomatic of a larger trend towards standardization in the mass culture industry
(2002[1941]: 47–48).

29. According to Ralf Hutter of Kra werk, the band began building its own drum machines in the mid-1970s because, ʻwe were C17.N29
really not pleased with the drum boxes you buy…we wanted more machine-like sounds because that is more what we are
aboutʼ (Aiken 1982).

30. Weheliye makes this argument as part of a larger intervention into post-humanist discourse, where he maintains that C17.N30
posthumanism, as commonly construed (e.g. in Hayles 1999), is implicitly White and fails to account for the exclusion of
Black subjects from Enlightenment-era accounts of what it means to be human (and thus, by extension, post-human).

31. A similarly nostalgic sentiment was recently expressed by veteran producer Brian Eno in a missive about how ʻwe can C17.N31
quantize everything now […but] of course, most of the records we like, all of us, as listeners, are records where people
didnʼt do everything to fix them up and make them perfectʼ (Spice 2015).

32. Incidentally, the MPC60 was also designed by Linn, a fact that extends his influence beyond disco, techno, and electro- C17.N32
pop, and secures his status as a household name amongst hip hop musicians. As Wu-Tangʼs RZA recently put it: ʻIf thereʼs
ever a hip hop hall of fame Roger Linn has to be inducted within the first year […] even to this day 80% of hip hop is
produced on that machineʼ (Noakes 2014).

33. The influence of J Dillaʼs o -kilter grooves has recently been addressed by Mike DʼErrico (2015), whose research on artists C17.N33
in the Los Angeles ʻbeat sceneʼ emphasizes their preference for taking beats ʻo the gridʼ via a mix of unquantized
sequencing and side-chain compression techniques. Likewise, Anne Danielsen has analysed what she describes as a
ʻseasick feelʼ in the music of DʼAngelo and other neo-soul artists (2010: 21).

34. Extending the argument, Scherzinger claims that ʻMIDI is a kind of colonialismʼ (2016: 62) due to the fact that it facilitates C17.N34
the imposition of Western musical values on non-Western cultures.
35. One such search for an alternative has been spearheaded by Xavier Serra, who edited a special issue of the Journal of New C17.N35
Music Research dedicated to ʻComputational Approaches to Art Music Traditions of India and Turkeyʼ (2014). In the issue,
several authors call for composition so ware to be built around the theoretical systems of Hindustani music from North
India or Makam music from Turkey.

36. Recent analytical work underscores this point. For instance, McGraw (2008) argues that western notation obscures the C17.N36
importance of fluid, fluctuating tempos in Gamelan music: ʻWestern theorists have quantized the temporality of Balinese
music into the literally repeated “cycle” ʼ (2008: 141).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
References C17.S7

Adorno, T. (2002[1941]). On popular music (with the assistance of George Simpson). In R. Leppert (ed.), Essays on music, 437–469. C17.P43
Berkeley: University of California Press.
Google Scholar Google Preview WorldCat COPAC

Agon, C., Assayag, G., Fineberg, J., and Rueda, C. (1994). Kant: A critique of pure quantification. Proceedings of the International C17.P44

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Computer Music Conference, 52–59. https://quod.lib.umich.edu/i/icmc/bbp2372.1994.015/—kant-a-critique-of-pure-
quantification?view=image.
Google Scholar WorldCat

Aiken, J. (1982). Kra werk: Electronic minstrels of the global village. Keyboard (Mar.): 33–40. C17.P45
Google Scholar WorldCat

p. 362 Aiken, J. (2000). Sequencers, part 5: Quantization. Keyboard 26(8): 122–124. C17.P46
Google Scholar WorldCat

Angliss, S. (2013). Mimics, menaces, or new musical horizons: Musiciansʼ attitudes toward the first commercial drum machines C17.P47
and samplers. In F. Weium and T. Boon (eds), Material Culture and Electronic Sound, 95–130. Smithsonian Institution Scholarly
Press.
Google Scholar Google Preview WorldCat COPAC

Auslander, P. (1998). Seeing is believing: Live performance and the discourse of authenticity in rock culture. Literature and C17.P48
Psychology 44(4): 1–26.
Google Scholar WorldCat

Battino, D., and Richards, K. (2005). Quantize this: Panel discussion. In D. Battino and K. Richards, The Art of Digital Music: 56 C17.P49
Visionary Artists and Insiders Reveal Their Creative Secrets, 133–138. Backbeat Books.
Google Scholar Google Preview WorldCat COPAC

Benadon, F. (2006). Slicing the beat: Jazz eighth-notes as expressive microrhythm. Ethnomusicology 50(1): 73–98. C17.P50
Google Scholar WorldCat

Berger, A. M. B. (2002). The evolution of rhythmic notation. In T. Christensen (ed.), The Cambridge History of Western Music C17.P51
Theory, 628–656. Cambridge University Press.
Google Scholar Google Preview WorldCat COPAC

Boenn, G. (2018). Computational models of rhythm and meter. Springer. C17.P52


Google Scholar Google Preview WorldCat COPAC

Brett, T. (2016). Virtual drumming. In R. Hartenberger (ed.), The Cambridge companion to percussion, 82–94. Cambridge C17.P53
University Press.
Google Scholar Google Preview WorldCat COPAC

Brøvig-Hanssen, R., and Danielsen, A. (2016). Digital signatures: The impact of digitization on popular music sound. MIT Press. C17.P54
Google Scholar Google Preview WorldCat COPAC

Brown, J., and Tucker, B. (1986). James Brown: The godfather of soul. Macmillan. C17.P55
Google Scholar Google Preview WorldCat COPAC

Buchanan, J. (2014). Understanding groove. Resident Advisor, 13 Aug. https://www.residentadvisor.net/features/2094 C17.P56


Google Scholar WorldCat

Butterfield, M. W. (2006). The power of anacrusis: Engendered feeling in groove-based musics. Music Theory Online 12(4). C17.P57
http://www.mtosmt.org/issues/mto.06.12.4/mto.06.12.4.butterfield.html.
Google Scholar WorldCat

Cemgil, A. T., Desain, P., and Kappen, B. (2000). Rhythm quantization for transcription. Computer Music Journal 24(2): 60–76. C17.P58
Google Scholar WorldCat

Clarke, E. (1987). Levels of structure in the organization of musical time. Contemporary Music Review 2: 211–238. C17.P59
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Cohen, T. F. (2009). The click track: The business of time. Metronomes, movie scores and Mickey Mousing. In G. Harper, C17.P60
R. Doughty, and J. Eisentraut (eds), Sound and music in film and visual media: A critical overview, 100–113. Continuum.
Google Scholar Google Preview WorldCat COPAC

Cont, A. (2008). ANTESCOFO: Anticipatory synchronization and control of interactive parameters in computer music. Proceedings C17.P61
of the International Computer Music Conference, 33–40.
Google Scholar Google Preview WorldCat COPAC

DʼErrico, M. (2015). O the grid: Instrumental hip-hop and experimentalism a er the Golden Age. In J. A. Williams (ed.), C17.P62
Cambridge companion to hip-hop, 280–291. Cambridge University Press.
Google Scholar Google Preview WorldCat COPAC

Dahlhaus, C. (1984). Die Musiktheorie im 18. und 19. Jahrhundert: Grundzüg einer Systematik. Wissenscha liche Buchgesellscha . C17.P63

Danielsen, A. (ed.) (2010). Musical rhythm in the age of digital reproduction. Ashgate. C17.P64
Google Scholar Google Preview WorldCat COPAC

Daston, L., and Galison, P. (2010). Objectivity. Zone Books. C17.P65


Google Scholar Google Preview WorldCat COPAC

Desain, P. and Honing, H. (1989). The quantization of musical time: A connectionist approach. Music Journal 13(3): 56–66. C17.P66
Google Scholar WorldCat

DeSantis, D., Hughes, M., Gallacher, I., Haywood, K., Knudsen, R., Behles, G., Rang, J., Henke, R., and Slama, T. (2016). Ableton C17.P67
reference manual version 9. Ableton AG.
Google Scholar Google Preview WorldCat COPAC

Diduck, R. (2017). Mad Skills: MIDI and music technology in the 20th century. Repeater Books. C17.P68
Google Scholar Google Preview WorldCat COPAC

p. 363 Do man, M. (2013). Groove: Temporality, awareness, and the feeling of entrainment in jazz. In M. Clayton, B. Dueck, and C17.P69
L. Leante (eds), Experience and Meaning in Music Performance, 62–85. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Eshun, K. (1998). More brilliant than the sun: Adventures in sonic fiction. Quartet Books. C17.P70
Google Scholar Google Preview WorldCat COPAC

Frith, S. (1986). Art versus technology: The strange case of popular music. Media, Culture, and Society 8: 263–279. C17.P71
Google Scholar WorldCat

Grant, R. M. (2014). Beating time and measuring music in the early modern era. Oxford University Press. C17.P72
Google Scholar Google Preview WorldCat COPAC

Hayles, K. N. (1999). How we became posthuman: Virtual bodies in cybernetics, literature, and informatics. University of Chicago C17.P73
Press.
Google Scholar Google Preview WorldCat COPAC
Hennig, H. (2014). Synchronization in human musical rhythms and mutually interacting complex systems. Proceedings of the C17.P74
National Academy of Sciences (9 Sept.): 12974–12979.
Google Scholar Google Preview WorldCat COPAC

Holden, J. (2015). Jamie Holden: On human timing. Ableton Live, 9 Mar. https://www.ableton.com/en/blog/james-holden- C17.P75
human-timing
WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Houghton, E. (2017). Modern approaches, hip-hop: Drums and percussion. Red Bull Music Academy Daily, 5 Oct. C17.P76
http://daily.redbullmusicacademy.com/2017/10/modern-approaches-hip-hop-drums.
Google Scholar WorldCat

Jackson, M. (2006). Harmonious triads: Physicists, musicians, and instrument makers in nineteenth-century Germany. MIT Press. C17.P77
Google Scholar Google Preview WorldCat COPAC

Keil, C. (1987). Participatory discrepancies and the power of music. Cultural Anthropology 2(3): 275–283. C17.P78
Google Scholar WorldCat

Kirnberger, J. P. (1982[1776–9]). The art of strict musical composition (trans. D. Beach and J. Thym). Yale University Press. C17.P79
Google Scholar Google Preview WorldCat COPAC

Langford, S. (2014). Digital audio editing. Focal Press. C17.P80


Google Scholar Google Preview WorldCat COPAC

Large, E. W. (2000). On synchronizing movements to music. Human Movement Science 19: 527–566. C17.P81
Google Scholar WorldCat

Lerdahl, F., and Jackendo , R. (1983). A generative theory of tonal music. MIT Press. C17.P82
Google Scholar Google Preview WorldCat COPAC

Linn, R. (1979). Modular drum generator. Filed on 1 Oct. 1979 in Hollywood, CA. https://patents.google.com/patent/US4305319. C17.P83
WorldCat

Linn, R. (1998). Perspective: Drumming with computers. Music and Computers 4(3): 22. C17.P84
Google Scholar WorldCat

Maelzel, J. (1815). Metronome or musical time-keeper. Filed on 5 Dec. 1815 in Middlesex, England. Repr. in Repertory of Arts, C17.P85
Manufactures, and Agricultures 33(2) (1818): 7–13.
Google Scholar WorldCat

Maelzel, J. (1822). Notice sur le Métronome de J. Malzel. Imprimerie de Constant-Chantpie. C17.P86

Manovich, L. (2001). The language of new media. MIT Press. C17.P87


Google Scholar Google Preview WorldCat COPAC

Mask, D. E. (2014). Looking for the perfect beat machine. Cuepoint, 4 Nov. https://medium.com/cuepoint/looking-for-the-perfect- C17.P88
beat-machine-3c110417f81a.
WorldCat

McGraw, A. C. (2008). Di erent temporalities: The time of Balinese gamelan. Yearbook for Traditional Music 40: 136–162. C17.P89
Google Scholar WorldCat

McNamee, D. (2009). Hey, whatʼs that sound? Linn Lm-1 Drum Computer and the Oberheim DMX. Guardian, 22 June. C17.P90
https://www.theguardian.com/music/2009/jun/22/linn-oberheim-drum-machines.
WorldCat
Mumford, L. (2010[1934]). Technics and civilization. University of Chicago Press. C17.P91
Google Scholar Google Preview WorldCat COPAC

Mynett, M. (2013). Contemporary metal music production. Doctoral dissertation, University of Huddersfield. C17.P92

Nero, I. (2004). Computers, cladding, and curves: The techno-morphism of Frank Gehryʼs Guggenheim Museum in Bilbao, Spain. C17.P93
Doctoral dissertation, Florida State University.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
p. 364 Newton, I. (1999[1687]). The principia: Mathematical principles of natural philosophy (trans. B. Cohen and A. Whitman). University C17.P94
of California Press.
Google Scholar Google Preview WorldCat COPAC

Noakes, T. (2014). Roger Linn: Doctor Beat. Dazed, 23 Aug. http://www.dazeddigital.com/artsandculture/article/21322/1/roger- C17.P95
linn-doctor-beat.
WorldCat

Ohriner, M. (2014). Listener–performer synchronicity in recorded performances of Chopinʼs mazurkas. Empirical Musicology C17.P96
Review 9(2): 101–128.
Google Scholar WorldCat

Pejrolo, A. (2011). Creative sequencing techniques for music production: A practical guide to Pro Tools, Logic, Digital Performer, and C17.P97
Cubase (2nd edn). Focal Press.
Google Scholar Google Preview WorldCat COPAC

Pierce, B., Oliver, M., Pierce, J. R., and Shannon, C. E. (1948). The philosophy of PMC. Proceedings of the Institute of Radio C17.P98
Engineers (Nov.): 1324–1331.
Google Scholar WorldCat

Questlove (2013). Conversation with Je ʻChairmanʼ Mao. Red Bull Academy Lecture Series. New York. C17.P99
http://www.redbullmusicacademy.com/lectures/questlove-new-york-2013.
Google Scholar Google Preview WorldCat COPAC

Repp, B. (1990). Patterns of expressive timing in performances of a Beethoven minuet by nineteen famous pianists. Journal of C17.P100
the Acoustic Society of America 88(2): 622–641.
Google Scholar WorldCat

Riskin, J. (2003). The defecating duck, or the ambiguous origins of artificial life. Critical Inquiry 29(4): 599–633. C17.P101
Google Scholar WorldCat

Roger Linn Design (2018). Past Products Museum. Accessed online 15 Aug. 2018. http://www.rogerlinndesign.com/past-products- C17.P102
museum.html.
WorldCat

Roland Corporation. (1980). CR-78 CompuRhythm Ownerʼs Manual. Hamamatsu. C17.P103


Google Scholar Google Preview WorldCat COPAC

Scarth, G., and Curry, O. (2013). DAW and drum machine swing. Attack Magazine, 1 July. C17.P104
https://www.attackmagazine.com/technique/passing-notes/daw-drum-machine-swing/.
Google Scholar WorldCat

Schedel, M. (2002). Anticipating interactivity: Henry Cowell and the Rhythmicon. Organised Sound 7(3): 247–254. C17.P105
Google Scholar WorldCat

Scherzinger, M. (2016). So ware physiognomics: Adornoʼs radio analytics today. New German Critique 43(3) (129): 53–72. C17.P106
Google Scholar WorldCat
Seeger, C. (1958). Prescriptive and descriptive music-writing. Music Quarterly 44(2): 184–195. C17.P107
Google Scholar WorldCat

Serra, X. (2014). Computational approaches to the art music traditions of India and Turkey. Journal of New Music Research 43(1): C17.P108
1–2.
Google Scholar WorldCat

Simondon, G. (2017 [1958]). On the mode of existence of technical objects (trans. C. Malaspina and J. Rogove). Minneapolis, MN: C17.P109

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470473 by National Science & Technology Library user on 26 May 2023
Univocal Publishing.
Google Scholar Google Preview WorldCat COPAC

Spice, A. (2015). The dangers of the digital: Brian Eno on technology and modern music. Vinyl Factory, 24 Feb. C17.P110
https://thevinylfactory.com/features/the-dangers-of-digital-brian-eno-on-technology-and-modern-music/.
WorldCat

Stanyek, J., and Piekut, B. (2010). Deadness: Technologies of the intermundane. Drama Review 54(1): 14–38. C17.P111
Google Scholar WorldCat

Sterne, J. (2003). The audible past: Cultural origins of sound reproduction. Duke University Press. C17.P112
Google Scholar Google Preview WorldCat COPAC

Théberge, P. (1997). Any sound you can imagine: Making music/consuming technology. University Press of New England. C17.P113
Google Scholar Google Preview WorldCat COPAC

Théberge, P. (2004). The network studio: Historical and technological paths to a new ideal in music making. Social Studies of C17.P114
Science 34(5): 759–781.
Google Scholar WorldCat

Todd, N. (1985). A model of expressive timing in tonal music. Music Perception 3(1): 33–57. C17.P115
Google Scholar WorldCat

Vail, Mark (1994). Rise of the machines. Roland CR-78, TR-808, and TR-909: Classic beat boxes. Keyboard 20(5): 82–89. C17.P116
Google Scholar WorldCat

Varèse, E. (1966). Liberation of sound (trans. C. Wen-Chung). Perspectives of New Music 5(1): 11–19. C17.P117
Google Scholar WorldCat

p. 365 Waters, R. H. (1948). Mechanomorphism: A new term for an old mode of thought. Psychological Review 55: 139–142. C17.P118
Google Scholar WorldCat

Weheliye, A. G. (2002). ʻFeeninʼ: Posthuman voices in contemporary Black popular music. Social Text 71(20): 21–47. C17.P119
WorldCat

Wiener, N. (1948). Cybernetics or control and communication in the animal and the machine. MIT Press. C17.P120
Google Scholar Google Preview WorldCat COPAC

Wilson, P. N. (2000). Vers une ʻécologie des sonsʼ: Partiels de Gérard Grisey et lʼesthétique du groupe de lʼItinéraire. Analyse C17.P121
e
Musicale 3 trimestre: 36–52.

Wilson, S. (2016). The 14 drum machines that shaped modern music. FACT Magazine, 22 Sept. C17.P122
http://www.factmag.com/2016/09/22/the-14-drum-machines-that-shaped-modern-music/
WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
CHAPTER

18 11-, 12-, and 13½-Bar Blues: Time and C18


African American
Country Blues Recordings (1925–1938) 
Andrew Bowsher

https://doi.org/10.1093/oxfordhb/9780190947279.013.21 Pages 366–C18.P129


Published: 08 December 2021

Abstract
This chapter examines commercially-issued recordings of African American country blues from the
early twentieth century, and considers the politics of representation involved with these recordings
related to the metric and structural orthodoxies of blues performance. Often featuring solo male
singers performing with guitar accompaniment, the recorded country blues of the 1920s–30s are
markedly exible in their approaches to timing. Drawing upon recordings of important country blues
artists including Blind Lemon Je erson, Robert Johnson, and Charley Patton, the chapter considers key
issues such as the controversy over the speed at which Johnson’s records were recorded, the exible
approach musicians took to the standard 12-bar format, and the strictures that the three-and-a-half
minute 78 rpm record side posed for artists’ songcraft. How these factors challenge musicological
orthodoxies over conventional blues structures and historical insights into the practice of the blues is
illuminated through the proposal that these recordings struggle with contentious narratives of
primitivism, racial stereotyping, and authenticity, whereby canonical 78 rpm records are rei ed to t a
prevailing narrative of the country blues as atavistic and authentic.

Keywords: blues, country blues, African American music, 12-bar blues, song structure, musical
performance, early sound recording, 78 rpm records, race, authenticity
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Introduction C18.S1

THE blues in the United States has historically been traced back to an encounter between W. C. Handy, the C18.P1
self-proclaimed ‘Father of the Blues’, and an anonymous singer on a railway platform in Tutwiler,
Mississippi, in the year of 1903. As the singer slid a knife over his guitar strings and sang about going ‘where
the Southern cross’ the Dog’, Handy recalled in his memoirs that it was ‘the most unusual music I have ever

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
heard’ (Evans 1982: 34). In time, Handy drew from this oft-mythologized encounter and started a
powerhouse commercial craze for the blues by composing such songs as ‘St Louis blues’ (1914) and ‘Yellow
dog blues’ (1915), which travelled across America on recordings made by White female singers and the
hugely successful sale of piano rolls and sheet music versions of these songs.

Jump forward 100 years. The wax cylinders that disseminated the rst recordings of Handy’s compositions C18.P2
have been succeeded by a myriad of music formats: 78s, 45s, LPs, cassette tapes, CDs, MP3s, and streaming
audio. One of the dominant popular music forms of the century— rock music—has had the blues’
ngerprints upon it since its inception in the 1950s. Louis Jordan, Fats Waller, and Chuck Berry were all
indelibly a ected by the blues craze that had had an enormous impact upon American music some 40 years
previously. Blues music itself then became a niche musical genre, tucked away in popular music record
shops next to other ‘speciality’ genres such as classical, reggae, jazz, and country.

p. 367 If you wish to nd music by the Father of the Blues in a record shop today, most likely you will encounter C18.P3
music not by Handy, but by a man hailed as the Father of the Delta blues: Charley Patton, a solo male
guitarist and singer who played country blues. From Papa Charlie Jackson’s emergence as the rst
commercially recorded solo male Black performer in 1925, through to the death of Robert Johnson in 1938,
recorded performers created the golden age of the country blues in the estimation of many consumers and
fans of the music (Dougan 2006; Hamilton 2007; Petrusich 2014; Wald 2004; Wardlow 1998). The singer is
most often a solo male performer, rarely accompanied by a second guitar, ddle, or female singer. His voice
is gru , intense, and pained. His acoustic guitar provides an idiosyncratic rhythmic thumping
accompaniment, full of slurred slide guitar lines and snapping bass strings. The sense of time is uid,
mercurial, and non-standardized to contemporary ears. His three-verse AAB blues verses might vary within
the same song from 11½ bars to 14½ bars. Yet, blues music is typically assumed to be built upon a strict 12-
bar form. The question that remains unaddressed is why pre-war country blues recordings seldom stick to
the orthodox 12-bar structure.

1
Minstrelsy, spirituals, vaudeville jazz, ‘classic’, and ‘country’ blues are genres that have become artefacts C18.P4
of cross-cultural ow, contact, representation, and appropriation. They persevere in the present day
through the recordings that were made in the eld by folklorists and commercial companies alike. The
recordings that were made have been commercially re-released ad nauseam, and the framing of these
recordings has undoubtedly a ected how the blues has been aestheticized and which aspects of blues
performance have been taken forward by progressive generations of listeners and practitioners. Music is ‘a
dynamic social text, a meaningful cultural practice, a cultural transaction, and a politically charged,
gendered, signifying discourse’ (Ramsey 2003: 18), and understood as such it is evident that current views
of blues practice and tradition are moulded by an array of interpretations that have been made of the music
during its history.

The ways in which Black American folk music forms were practised, sustained, and appraised changed C18.P5
dramatically during the twentieth century: the recording industry and early ethnographic eld recordings
have largely informed the common perception of Black musical forms. To paraphrase Marybeth Hamilton
(2007), these recordings are Black voices altered to t a White vision of Blackness, which has been largely
informed by those middle-class White collectors who ‘discovered’ the blues in the 1950s and gave rise to the
large White-oriented market for these recordings and to the canonization of the artists who created them
(Dougan 2006). In essence, emic folk categories and preferences have been blurred and recast according to
White analytical etic categories and the marketing categories of record industries, which have reshaped
perceptions of African American country blues.

It is ironic that as early as 1939 Alain Locke was bemoaning that ‘one of the handicaps of Negro music today C18.P6
is that it is too popular […It is] tarnished with commercialism and the dust of the marketplace’ (Locke 1939:
4). The creation of the current mainstream understanding of ‘authentic’ blues and Black folk music still

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
p. 368 betrays the in uence of the architects of the blues and folk music revivals of the 1950s and 1960s (Hamilton
2007; Skinner 2006; Weissman 2006). They wed race, ethnicity, space, and music together in a way that
is still exploited by businesses dealing in the production and sale of commodi ed representations of identity
(Dougan 2006; Pearson 1992).

Time, the blues, and the politics of representation C18.S2

This chapter interrogates how notions of performance and musical time are realized in prewar recordings of C18.P7
country blues. Using a series of case studies I examine how metrical variations occur across di erent
recordings of the same song by the same artist (Blind Lemon Je erson’s ‘Matchbox Blues’, 1927) illustrates
these performative variants and how the performance of metre and song structures in the country blues
varies across the recorded performance of a single song (highlighted by Charley Patton’s ‘Pony blues’,
1929). I then explore the uidity of tempo in blues recordings, and how recording and the technology of the
1920s and 1930s have a ected the understanding of canonical blues song structure and performance
(discussed with recourse to Sam Collins’s recorded repertoire, as well as the recordings of Je erson, Patton,
and Robert Johnson). Finally, I consider how revisionist, atavistic, and primitivist narratives of the blues jar
with the recorded works of country blues’ most famous exponent, Robert Johnson.

First, I frame these case studies by discussing blues and the politics of representation. Those who control C18.P8
the popular narrative of the blues have largely been White—from the talent scouts, record labels, and
recording engineers and producers in the country blues heyday (Wardlow 1998: 126–151 on H. C. Speir is
especially instructive) to those who started collecting blues recordings and began writing on the subject
from the 1940s onwards. This has resulted in a distinct ‘othering’ of the blues and its progenitors. Johannes
Fabian (2002) has suggested that anthropologists create their subject by metonymically freezing them in
time, denying them histories or a state of progress, essentially distancing them from a world that is self-
consciously historical and seeking a future. It might be argued that consumers of the blues have similarly
frozen country blues artists in time, along with their geographical landscapes of rural poverty and proximity
to the Mississippi Delta. Time and place have been fetishized as ahistorical, creating a fallacy of the rural
Deep South and imposing an aural landscape upon the people according to a particular canon of 78 rpm
recordings. The gendering and racial segregation of blues artists has been similarly e ected in a way that
highlights the perceived archetype of the blues performer over other performers and genres in the African
American traditions of the time. Thus, the history of the blues has largely been written according to the
demands, whims, and agendas of primarily White gatekeepers who have marginalized multiple iterations of
African American performance whilst giving primacy to a particular idealized trope of blues practice.

p. 369 These gatekeepers—among them the famed early ‘Blues Ma a’ of James McKune, Pete Whelan, Harry C18.P9
Smith, and others (Hamilton 2007; Petrusich 2014: 127–129)—practised a sort of romantic obscurantism
that fetishized records for their rarity. Records by Leroy Carr, Blind Blake, and Blind Lemon Je erson might
have sold hundreds of thousands of copies, but those by less prodigious sellers such as Tommy Johnson,
Willie Brown, King Solomon Hill, Robert Johnson, and Son House became more sought after and
subsequently dei ed in the blues canon. Temporality is therefore manifested in the curation of cultural
history by collectors who developed the pantheon of great blues artists and have helped to shape
perceptions of a recognized canon (Dougan 2006; Gioia 2008; Petrusich 2014; Skinner 2006).

The durability or transience of material culture is often socially produced. Those controlling consumer C18.P10
tastes and the artefacts that feed their desires can keep the blues alive for and with a White audience even if
2
the Black community does not consume the music. Subsequently, the type of blues that is passed on, and
the music recorded contemporaneously but neglected, is decided upon by those who control social

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
durability. These forces can sell millions of Robert Johnson albums—resurrecting the career of a long-dead
musician who was unknown in his lifetime—through promoting the mythological allure of the Delta blues
musician who purportedly sold his soul to the Devil (Gioia 2008: 160–169) and died a violent, premature
death; yet they marginalize in uential artists of his time such as the Mississippi Sheikhs and Leroy Carr,
who sold hundreds of thousands of records to African Americans in the late 1920s. Listening, like looking at
paintings, might be seen to be a ‘privilege of domination’ in this context (Myers 1995: 45).

Blues genres and traditions C18.S3

In order to elucidate why time and metre are so contentious in country blues performances and recordings, C18.P11
it is necessary rst to contextualize country blues relative to other blues styles of the 1920s and 1930s. The
‘classic’ blues has its origins in the career of W. C. Handy, and is intimately entwined with the music
marketplace of the 1910s–1920s. Handy’s compositions and subsequent classic blues recordings were
hugely popular, and initially featured female singers fronting small orchestras and hot jazz bands, whilst
also betraying the in uence of America’s nineteenth-century vaudeville and minstrelsy traditions (Lott
1993). The rst commercial releases were of White singers, until OKEH Records released Mamie Smith and
the Jazz Hounds’ ‘Crazy blues’ in 1920. The record proved to be a huge commercial success, as many African
Americans connected with recordings that featured other Black Americans. This led to the emergence of the
‘Race Records’ industry, with several labels signing Black female artists and releasing recordings by such
singers as Ida Cox, Bessie Smith, and Ma Rainey. These artists and their repertoires were enormously
in uential, and they toured the country to large crowds playing sophisticated revue-style concerts that
highlighted the artists’ urbanity and glamour (Davis 1998; McGinley 2014). The market for Race Records
p. 370 began in the 1920s, and ourished among Black audiences until the Depression hit America’s record
industry in the early 1930s. It produced thousands of recordings of blues, gospel, string bands (remarkable
in their similarity to minstrel groups and White country), and sermons.

That early classic blues recordings almost uniformly conform to standard 8-, 12-, and 16-bar structures is C18.P12
largely due to the fact that these are ensemble performances where a lack of adherence to strict bars and bar
lines could result in chaos. Furthermore, these songs were often written by professional songwriters and
disseminated by sheet music and piano rolls which speci ed these rigid performance structures. The
country blues was not, however, the product of such industry forces, but rather represents a folk music
tradition that was not as lucrative as the classic blues. Ensembles in the country blues are notably smaller
than those in classic blues recordings, and the majority of such recordings were made by solo artists. That a
larger ensemble would not be as feasible a commercial proposition is perhaps of signi cance to this fact.
Also, country blues songs were learned in an oral tradition (Evans 1982: 253–255) in which local guitar
styles and lyrical verses were adopted and adapted by many musicians without the use of sheet music or
piano rolls to mediate their performances. When 78 rpm records started to spread performance styles and
lyrics, these still acted as a template upon which other musicians would extemporize rather than copy
3
verbatim. That country blues songs were not transmitted or borne from metrically precise musical notation
but through performance is a likely contributor to the genre’s characteristic metric uidity.
Male African American blues singers were not recorded commercially until Papa Charlie Jackson in 1926, C18.P13
and his solo banjola-led performances proved successful enough for scouts and labels to expand their
e orts to capitalize on the African American rural marketplace. This came at a time when many record
labels were turning their attention to rural audiences and ethnically conceived demographics that
previously had not been harnessed or catered for. Therefore, there had been releases of White folk music by
Fiddlin’ John Carson in 1923, Cajun music from Louisiana, recordings of Jewish cantors, and many other
niche genres (Kuhn 2005; Peterson 1997: 12–32) before the country blues was recorded commercially. At the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
same time, along with these rural communities, many African Americans and minority groups were living in
America’s cities (Keil 1966), where their styles were changing in response to both the formalized music of
the urban environment and their isolation from rural audiences. Urban blues’ breakthrough male star was
Leroy Carr, a pianist and singer (often accompanied by Scrapper Blackwell’s guitar) who recorded
extensively in Chicago prior to his premature death in 1935. Carr’s music is suave and rhythmically precise,
and made great use of the electronic microphone which enabled him to croon softly in a style quite the
4
opposite of the belting female vocal delivery of ‘classic’ blues recordings. Carr was commercially successful
and very in uential, with many of his tunes becoming standards and entering the repertoires of rural
artists. The corpus of songs for which he is best remembered—including ‘How long, how long blues’ (1928),
‘How long has that evening train been gone’ (1932), and ‘Blues before sunrise’ (1932)—do not represent the
full extent of his repertoire. Like many solo blues musicians, Carr was a rounded entertainer, and his
compositions include such novelty songs as ‘Papa wants a cookie’ (1930) and ‘Carried water for the
p. 371
elephant’ (1931). In the presentation of Carr’s artistry today (as with many other blues legends of the 1920s
and 1930s), the latter songs are often left unaccounted for in favour of his blues material, which mostly
follows the classic 12-bar structure. Carr and his contemporaries are often thought of as ‘slick’ and
‘inauthentic’, a long-standing aesthetic prejudice that dates back to the earliest creation of the blues canon
by White collectors despite the popularity and success of Carr and his contemporaries (Wald 2004: 237–
240).

Another consideration of where country blues ts within the contemporary agenda of blues appreciation is C18.P14
the e ect that eld recordings have had upon the legacy of African American music of the 1920s–1930s.
These recordings, and the publicity accorded to the presentation of artists such as Leadbelly (famously
featured in Life magazine’s 19 April 1937 edition under the title ‘Bad nigger makes good minstrel’—see Wald
2004: 232) and the principal players of the blues revival (‘rediscovered’ artists such as Mississippi Fred
McDowell, John Hurt, Son House, and Skip James), point to blues as being a music of the disenfranchised
and those on the cultural margins of even the Jim Crow-era African American experience. This viewpoint is
compounded by the recordings made by Alan and John Lomax at Parchman Farm State Penitentiary and
5
other prisons in the USA. The narrative is a familiar one of salvage ethnography and the goal of
documenting the ‘purest’ culture possible—that which is on the margins of society and is therefore
unfettered by the demands and whims of commercial marketplaces. Even today, ‘real’ country blues is seen
6
as a music of the margins only available from African Americans in the rural backwaters of Mississippi.
This revisionist narrative created a binary opposition between the marketplace for commercial blues
recordings by professionals such as Carr which are ‘impure’, and those recorded by isolated amateurs such
as the inmates of jails where the culture is unsullied and ‘pure’. The irony of course is that John Lomax
lobbied for and invested heavily in Leadbelly’s post-Parchman stardom (Szwed 2011: 59–76), and that,
conversely, many of the people who recorded commercial country blues were themselves amateur artists
who often recorded opportunistically or for just a single session. These performers were also often
in uenced by commercially released music, and modelled their songs and performances after artists who
had achieved some fame and notoriety as recording stars (evidenced by the longevity of Blind Lemon
Je erson’s ‘Matchbox blues’ and Robert Johnson’s ‘I believe I’ll dust my broom’ (1936), to name but two
examples). But in the climate of the postwar blues ‘ma a’ (Hamilton 2007; Petrusich 2014: 127–129), the
homespun and rural became dei ed over the commercially orchestrated and urban. Classic blues became
commercial fodder, whereas country blues was considered ‘the real deal’ (Wald 2004: 220–249).
Country blues C18.S4

In this entangled web of blues genres, the country blues stands today as the most revered of the prewar C18.P15
p. 372 blues styles. A cult of authenticity has arisen over the artists, their songs, their geographical
backgrounds, and their record labels that paints a revisionist history of the blues landscape in the early
twentieth century. The most revered of these performers are solo performers whose music is deemed to

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
have been born out of the estrangement and disenfranchised position of rural Black America. The blues is a
personal music re ecting life’s hardships and the speci c predicament of those on the fringes of society
(Ottenheimer 1979)—such is the central dictum that guides such in uential studies of the blues as Paul
Oliver’s Blues fell this morning (1960) and Samuel Charters’ The country blues (1959).

From the extant recordings and anecdotes about these musicians, they did not just perform blues, but also C18.P16
7
folk ballads (often sharing their repertoire with that of the White recording artists of the time), gospel,
instrumentals, and ragtime songs. They were, in the common parlance, ‘songsters’ as much as they were
‘bluesmen’ (Oliver 1984). This was especially true for those who both performed in churches and acted as
part-time preachers, and those who performed at dances (often for White patrons) where a varied
repertoire was required. Street singers also had to harness a wide variety of material in order to earn a living.

Many of the blues artists of this period have become synonymous with their geographical origin, such as C18.P17
8
Texas, the Piedmont, or the Mississippi Delta. Whilst the latter region has become enshrined as the cradle
of the blues’ most authentic and interesting styles and performers, the rst major star of the country blues
was Texas’s Blind Lemon Je erson, a skilled guitarist and singer who recorded 94 sides in 1926–1929, and
was in uenced by American and Mexican vernaculars (Narváez 2002). Just as Appadurai (1988) has
highlighted the ‘problem of place’ and the need to resist topological stereotypes which lead ethnographic
places to become metonyms for anthropological ideas, this localizing view of the blues needs to be
interrogated. If the production of a sense of locality—such as a site of Black folk-authenticity made by
White subjects—is achieved through a ritual localizing and spatializing of time and space, the creation of
records and the imagery used in packaging and advertising these recordings is a contributing factor in
creating a site of Blackness and a repository for blues’ authenticity. African American authenticity becomes
cosseted into the past, into the poor rural backwaters of the American South where the most ‘primitive’
(and therefore most ‘authentic’) practitioners of Black music dwell in the eyes of White consumers
(Hamilton 2007: 15–22). Cultures often do not exist within strict boundaries and locations, yet there is a
touristic ‘fetish of locality’ (Cook and Crang 1996: 136) evident in popular conceptions of blues, whereby
‘consumed commodities and their valuations are divorced for and by consumers of their production and
provision through the construction of ignorance about the biographies and geographies of what we
consume’ (p. 135). The fetish is the construct of modern White consumerism and the narrative of twentieth-
century popular music, of the consumption of the material goods (music) and its active engagement in
social self-creation, and the creation of networks and groups linked through this common identity or
preference. This recontextualization of goods is not just a product of their possession and consumption by
White people; it is historically informed, harbouring ickers of older notions of romantic folk societies and
p. 373 the exotic ‘other’ (cf. Aubert 2007). Nevertheless, the archetypical country bluesman is perceived as an
impoverished African American male of rural origin who sings and accompanies himself on guitar. This
image persists despite the popularity of piano blues, jug bands, and string bands in the vernacular music
traditions of the southern United States, and the uidity between musical genres and racial barriers that the
music evidences.
The standard blues forms C18.S5

The standard call at many a jam session is to ‘play a blues’. This almost invariably means a 12-bar structure C18.P18
based upon the tonic, subdominant, and dominant of the scale, although the exact performance of this
might vary as follows:

Example A: C18.P19

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
I IV (or I I) C18.P20

IV IV I I C18.P21

V IV I I C18.P22

Example B: C18.P23

7 7
I IV I I C18.P24

7 7
IV IV I I C18.P25

7 7 7
V IV I/IV I/V C18.P26

Example B shows one of numerous possible variations involving chord extensions and a half cadence (or C18.P27
turnaround). This variation, like others, clearly relates to the basic model of Example A. Vocal
accompaniment to these 12-bar forms usually takes an AAB form:

My mama’s dead and my papa can’t be found C18.P28


My mother’s dead and my papa can’t be found C18.P29
I ain’t got me nobody to throw my arms around C18.P30

Sam Collins, ʻDevil in the lionʼs denʼ (1927) C18.P31

Other iterations are found in prewar blues structures, based again on the tonic, subdominant, and dominant C18.P32
of the key. These may take a 16-bar form, with a verse structure of AAAB. In such a case, the chord structure
of the second line is repeated. There are also a variety of variations on 8-bar musical structures, with the
verse form AB:

I know my dog anywhere I hear him bark C18.P33


Well I’d know my rider if I seen her in the dark C18.P34

Blind Boy Fuller, ʻPistol snapper bluesʼ (1938) C18.P35

p. 374 Such an 8-bar blues might take the form of Example C: C18.P36

Example C: C18.P37

I V I IV C18.P38

7 7
I/VI II/V I I C18.P39

In such contexts as performing in a rock jam session or playing a hard-bop blues, not being able to adhere C18.P40
strictly to these changes would be cause for being considered ‘unable’ to play the blues. Solo performers
conventionally follow the metrically precise structures outlined above. Such postwar solo guitarists and
singers as Taj Mahal, Eric Bibb, Keb’ Mo’, Ste an Grossman, John Fahey, Bert Jansch, John Hammond Jnr,
C. W. Stoneking, and Duck Baker have routinely performed blues according to the strictures of the
conventional form. The repertoires of all these performers draw from the country blues recordings that are
analysed in this chapter. What is striking is that these revivalist postwar performers pay rigid respect to
formal structural conventions rather than emulating their in uences in displaying such high-wire musical
exibility as to play 11-, 13-, or 14¼-bar choruses.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
Time and the structure of the blues: Four case studies C18.S6

With this historical context in mind, I now turn my attention to four case studies involving country blues C18.P41
artists of the 1920s–1930s that demonstrate the genre’s exibility of timing in musical performance, and
how these idiosyncrasies relate to the narrative of revisionist blues histories outlined above.

Blind Lemon Je erson C18.S7

As mentioned above, the rst star of the country blues is generally considered to be Blind Lemon Je erson. C18.P42
He was born in the town of Couchman in 1897, but was principally active in Dallas. He was proli cally
documented from 1926 until his death in 1929 from myocarditis, recording 94 sides of which many sold in
six gures. Je erson started out as a street singer—a not-uncommon career choice for blind African
Americans at the time—and his high-pitched voice and hard attack on the guitar lent themselves well to
being captured and reproduced via early recording technologies. The signi cance of Je erson and his
success lies in the fact that his commercial clout at Paramount Records opened the doors for other solo male
blues guitarist/singers to enter recording studios. Many of his songs became standards not only amongst his
contemporaries but also within the folk revival, and eventually his ‘Matchbox blues’ became a rock ’n ’roll
favourite in the hands of Carl Perkins and The Beatles.

p. 375 Like many of the early country blues artists, Je erson had a small pool of largely rhythmic ‘stock’ guitar C18.P43
arrangements that he used for the majority of his compositions. His familiarity with these arrangements
would have enabled him to improvise his guitar lls with considerable freedom, con dent that he would not
forge himself a musical riddle out of which he could not nd a way. Indeed, Je erson took a great deal of
liberty with the timing of his stock arrangements, allowing out-of-metre interjections and single-string
runs to break the steady progress of a 12-bar blues sequence. His guitar playing is muscular but loose-
limbed, and unencumbered by an accompanist for the majority of his recordings (barring three sides cut
with George Perkins’s piano in Chicago in May 1927) and his deft clusters of improvised guitar lines slash
across the bar lines or hang suspended whilst Je erson elongates a line in his vocal delivery.

‘Matchbox blues’, one of Je erson’s signature songs, was recorded on three separate occasions for the C18.P44
Paramount and Okeh labels between March and April 1927. The song follows the 12-bar structure that
Je erson used for many of his compositions (the exceptions being his version of the folk ballads ‘Jack
O’Diamond blues’, 1926, and ‘Beggin’ back’, 1926, the ragtime-in uenced ‘Hot dogs’, 1927, his recordings
of religious materials, and the folk-revival standard ‘See that my grave is kept clean’, 1927 and 1928), but
the text for ‘Matchbox blues’ varies considerably across these three performances. As David Evans observes
(1982: 78):

[t]here are only two stanzas that are used in all three versions, whereas each of these versions has C18.P45
six or seven stanzas altogether. One other stanza is used in two of the three versions, but otherwise
all of the other eleven stanzas he sings are di erent. It would appear, then, that only two stanzas
and the melody and the guitar part formed a stable unit in Je erson’s repertoire, while the rest of
the stanzas were added at the time of performance.
It is certainly conceivable that in the context of live performance all these stanzas (and possibly more) could C18.P46
have featured in extended performances of the song unencumbered by the strictures of recording a 78 rpm
9
record side. The practice of splitting songs over two sides of a 78 rpm record in a Part 1/Part 2 structure
indicates that blues songs were often performed for much longer than the three-minute side could
accommodate. Recordings of artists such as Skip James, Son House, Fred McDowell, and Bukka White from
the 1960s further attest to the potential of blues songs to last much longer than a 78 rpm side would allow.
Bukka White’s ‘Sky songs’ (1965) at times unfolds over 15 minutes—a radical departure from the 3½-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
minute concision of the 78 rpm records he produced in the 1930s.

When the metric treatment of the song is analysed, it becomes clear that Je erson takes liberties with the C18.P47
12-bar sequence di erently in each iteration of ‘Matchbox blues’. That this should happen in three versions
recorded over a short two-month period pinpoints the essential elasticity of musical time that was a crucial
p. 376 aspect of Je erson’s musical practice. In the following analysis, I examine two stanzas that are repeated
in all three versions of ‘Matchbox blues’ (with some subtle variations—see Titon 1994: 34–36):

Stanza A C18.P48

Standin’ here wonderin’ will a matchbox hold my clothes C18.P49


I was settin’ here wonderin’ will a matchbox hold my clothes C18.P50
I ain’t got so many matches but I got so far to go C18.P51

Stanza B C18.P52

Lord, mama, who may your manager be C18.P53


Hey hey mama, who may your manager be C18.P54
Reason I ask so many questions C18.P55
Can’t you make ’rangements for me C18.P56

(Transcribed by Titon 1994: 34–35) C18.P57

The stanzas appear in di erent places in each of the performances (see Table 18.1). C18.P58

Table 18.1 Comparison of stanza placements in three separate recordings of ʻMatchbox bluesʼ C18.T1

Version of ʻMatchbox bluesʼ No. of stanzas Position of stanza A Position of stanza B

1 6 2 3

2 6 1 6

3 7 1 2

10
Each of these three recordings of ‘Matchbox blues’ employs a di erent metric structure for the I-IV-V 12- C18.P59
bar blues arrangement of both stanzas. The rhythmic and metric variation of Je erson’s playing is such that
it resists transcription using sta notation. For this reason, Tables 18.2 and 18.3 provide approximate
structures in bar lengths of stanzas A and B across the three recordings.

The improvisations that Je erson performs on the guitar are sometimes familiar from take to take, with C18.P60
both versions 2 and 3 of stanza B containing a rubato, almost free-time ostinato during the delivery of the
second line of the stanza, but the metric parameters in which these improvisations unfold are ambiguous
and constantly shifting. To complicate matters further, while loosely aligned, Je erson’s vocal line and the
guitar accompaniment are sometimes metrically displaced, with the guitar often joining the vocals after a
short delay. In version 1 of the second line of stanza A, for example, the vocal line resolves onto I from IV
several beats ahead of the guitar. Moreover, while the bar lengths presented here are based on a
conventional 12-bar blues time signature of 4/4, there are several instances where the sense of a regular
metric structure is lost completely. This is evident in version 3 of rst two lines of stanza B, where the
tremolo gures in the guitar undermine the music’s pulse so that there is no discernible metre. This means
p. 377 that the 17 crotchet beats of line 1 might be subdivided as four 4/4 bars plus one crotchet beat, or perhaps

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
a combination of 3/4, 4/4, or 5/4 bars, depending how the rubato applied to particular beats is interpreted. A
similar disruption of metre is created in version 1 of stanza B by the use of fermatas on ‘Lord’, ‘mama’ in the
rst line, and ‘Hey’ in the second line, accompanied by guitar tremolo. This instability is further
compounded by the harmonic changes employed by the guitar, which sometimes depart from the standard
I-IV-V structure. In version 2 of stanza A, for instance, the rst four beats of line three appear to substitute
chord ii for chord V, followed by six beats of chord V and ten beats of chord I. This exible switching
p. 378 between chords ii and V also occurs frequently in version 3. Je erson stretches or condenses various
musical features within a phrase to create the varying phrase lengths. In version 1 of stanza A, the longer
length is caused by a fermata on ‘here’ and the extended nger-plucked ri at the end of the phrase;
extended guitar ri s are used similarly to extend the subsequent two phrases. Elsewhere (as at the end of
the third line of stanza A, version 2), the length of a phrase might be extended through the repetition of the
nal chord in the guitar. This rhythmic and metric elasticity is present throughout all three versions of the
song. It should be noted that contemporary players interviewed in later years did not mention metric
instability in discussions of Je erson’s playing. He was almost uniformly lauded by those who recalled his
playing, and such was Je erson’s stature that he was gifted a number of eulogies by fellow recording artists
such as Rev. Emmett Dickenson’s ‘Death of Blind Lemon’ (1930) and ‘My friend Blind Lemon’ (1935) by
Leadbelly (see Lornell, 2000).

Table 18.2 Structure of stanza A across three recordings of ʻMatchbox bluesʼ C18.T2

Version Chord Line Length (in bars)

1 I Standinʼ here wonderinʼ will a matchbox hold my clothes 6

IV I was settinʼ here wonderinʼ will a matchbox hold my clothes 6½

V I ainʼt got so many matches but I got so far to go 4¾

2 I Iʼm settinʼ here wonderinʼ will a matchbox hold my clothes 5

IV Iʼm settinʼ here wonderinʼ will a matchbox hold my clothes 5½

V I ainʼt got so many matches but I got so far to go 5½

3 I Iʼm settinʼ here wonderinʼ will a matchbox hold my clothes 5½

IV Iʼm settinʼ here wonderinʼ will a matchbox hold my clothes 5

V I ainʼt got so many matches but I got so far to go 4¾


Table 18.3 Structure of stanza B across three recordings of ʻMatchbox bluesʼ C18.T3

Version Chord Line Length (in bars)

1 I Lord mama who may your manager be 3½

IV Hey hey mama who may your manager be 5½

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
V Reason I ask so many questions canʼt you make ʼrangements for me 4½

2 I Now tell me mama who may your manager be 6

IV Now tell me who may your manager be 5¾

V Reason I ask so many questions canʼt you make ʼrangements for me 5

3 I I said fair brown who may your manager be 4¼

IV Oh mama who may your manager be 5

V Reason I ask so many questions canʼt you make ʼrangements for me 5

It is important to take note of the opinions of blues recording artists’ peers. The country blues artists were C18.P61
not seen as sloppy or inept for not replicating the crisp rhythmic regularity and sheet-music polish of the
classic blues singers. Indeed, it is arguable that their popularity lay in their presentation of a form of
localized blues that sounded familiar to the record-buying African American public. Performers such as
Je erson, for all their wayward rhythmic conceptions, probably treated musical time in the same way as
guitar-playing Black musicians around the country (as evidenced by recordings made both for commercial
labels and for ethnographic research). The imprint of the classic blues is to be found in the ‘ oating
11
verses’ that country blues took from its urban counterpart, as well as such experiments as Je erson’s
‘Rising high water blues’ (1927) in which he sings over the metrically regular piano accompaniment of
George Perkins. However, in the solo context where singer and guitar dictate the ow of a performance, the
metre of country blues tends to have a more elastic conception compared with the strictures of the classic
12-bar blues.

Charley Patton C18.S8

Charley Patton is widely regarded as one of the most signi cant of the early rural blues performers. Indeed, C18.P62
12
one of the rst compilation releases to feature Patton’s music exclusively was on Yazoo entitled Founder of
the Delta Blues (1971). Patton is seen as a mentor to many of the other signi cant Mississippi bluesmen,
including Muddy Waters, Son House, and Robert Johnson. Patton is even credited with having taught such
signi cant artists as Willie Brown and Howlin’ Wolf (who shares Patton’s gru vocal intensity) how to play
the guitar.

Patton was largely a solo performer, and as a guitarist he possessed a loud and rhythmic style ideally suited C18.P63
to the dances that he accompanied as a professional musician. He was famed for thumping a consistent
pulse on the guitar and stamping the oor to add to the percussiveness of his playing, and was a
consummate entertainer who played the guitar behind his back, under his leg, and by tossing it in the air. He
p. 379 was a Mississippi native who lived on the Dockery Plantation, entertaining locals and itinerant workers
as well as being in high demand for dances hosted by both Black and White patrons. Patton led a
tempestuous life blighted by alcohol and violence, including an incident in 1933 that left him with a large
knife-scar along his throat.
But hereafter, his correspondence with the country blues archetype becomes problematic. Patton was a C18.P64
successful performer who prided himself on driving new cars and being impeccably attired. Indeed, the only
known photo of Patton shows him wearing a bow tie, pinstriped two-piece suit, and spats, with his collar
slightly raised on the right, as if to hide his scar. Furthermore, Patton is clearly mixed-race, and
contemporaries such as Howlin’ Wolf attest to him being of mixed Cherokee, Mexican, White, and Black
heritage. His repertoire also runs the gamut of ragtime, gospel, folk ballads, and hokum songs in addition to
the blues. In his study of Patton’s music, John Fahey (1970) identi es 33 out of 52 of Patton’s then-known

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
recordings as blues. He also transcribes 22 of Patton’s blues stanzas, and nds that only 8 accord to the
classic 12-bar structure, with others consisting of 11½, 12¼, 12½, 13, 13½, 14, and 14½ bars. Whilst the
harmonic progression of the 12-bar blues underpins these performances, Patton displays an unerring sense
for musical drama in not sticking to rigid metric conventions. A ne example is ‘Moon going down’.
Recorded in Grafton, Wisconsin, in June 1930, the performance features a rare duet with Patton’s protégé
Willie Brown on second guitar. What is signi cant about this performance is that, despite the constraints of
playing with an accompanist, Patton directs the performance to deviate from the 12-bar form, but still
resolving into a consistent non-standard metric pattern. Part of this is indicated in Patton’s vocal, where he
repeats phrases to elongate the duration of phrases. This is signi cant in the fourth and fth stanzas (see
Table 18.4).

This approach results in two verses that are identical in their length and metric structure (14 ½ bars). That C18.P65
these two stanzas that deploy the repetition of elements of the rst line have complementary structures
suggests that this was a de nite rhythmic device which Patton would deploy, yet other stanzas follow
di erent structures (see Table 18.5).

Patton’s accompanist, Willie Brown, is not only nimble in his adherence to these imprecise musical C18.P66
changes, but also exhibits similar exibilities on his extant contemporaneous recordings (‘MandO blues’
(1930) and ‘Future blues’ (1930)). This would suggest that Patton’s timing is far from atypical of the country
blues performer, but was instead a conventional rhythmical conception amongst musicians of the
Mississippi Delta.

Another of Patton’s 12-bar records, ‘Pony blues’, inspired many spin-o records from other artists. C18.P67
Recorded at Patton’s rst session in Richmond, Indiana, on 14 June 1929, ‘Pony blues’ was a commercial
success and led to a famous Paramount Records advertisement (Dahl 2016: 111). Though it depended upon
13
the local ‘Jinx/Maggie Campbell’ pattern for the majority of its stanzas, ‘Pony Blues’ nonetheless displays
a range of rhythmic devices and o -kilter approaches to the 12-bar form (see Table 18.6).

For the stanzas II, III, and V, Patton plays a gure high up on the guitar that accompanies the three-bar C18.P68
duration of the rst line in each stanza. It therefore seems that this guitar gure was consistently
performed in a way that dictated the metric qualities of Patton’s performance. The 12-bar blues can
therefore be broken down into a harmonic progression that allows rhythmic licence to performers, where
p. 380 particular lines and guitar gures that might disrupt the regularity of 12-bar blues can be assimilated
into a high-wire act of maintaining a regular pulse whilst simultaneously being exible with the music’s
metric structure.
Table 18.4 Comparison of line lengths in Charley Pattonʼs ʻMoon going downʼ C18.T4

Stanza Line Length (in bars)

IV Lord, I think I heard that Helena whistle, Helena whistle, 5½

Helena whistle blow

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
Lord, I think I heard that Helena whistle… 4½

[Spoken] Well I hear it now

Lord, I ainʼt gonna stop walkinʼ tilʼ I get to my riderʼs door. 4½

V Lord the smokestack is black and the bales they shine like, bales they shine like, 5½

bales they shine like gold.

The smokestack is black and the bales they shine like gold. 4½

[Spoken] Well yʼknow, boy it sure look good to me!

Lord I ainʼt gonna walk the levee ʼround no more. 4½

Table 18.5 Comparison of stanza lengths in Charley Pattonʼs ʻMoon going downʼ C18.T5

Song Length of tonic line 1 (in Length of subdominant line 2 Length of dominant line 3 Total no. of
section bars) (in bars) (in bars) bars

Intro 1½ N/A N/A 1½

Stanza I 5½ 4½ 4½ 14 ½

Stanza II 4½ 4½ 4½ 13 ½

Stanza III 4 4½ 4½ 13

Stanza IV 5½ 4½ 4½ 14 ½

Stanza V 5½ 4½ 4½ 14 ½

Stanza VI 4 4½ 3½ 12
Sam Collins C18.S9

Sam Collins, born in 1887 in Pike County, Mississippi, was one of the oldest blues practitioners to have been C18.P69
recorded by commercial labels in the 1920s and 1930s. His songs recorded between 1927 and 1931 range from
the rst commercial recording of the folk ballad ‘Midnight special’ (1927) and the rag of ‘Salty dog’ (1931),
14
p. 381 to the gospel of ‘I want to be like Jesus in my heart’ (1927). Collins’s main reputation rests on his blues
recordings, showing his high keening vocal over mostly ngerpicked solo guitar in an open minor tuning,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
and featuring liberal doses of slide guitar. Such songs as ‘Devil in the lion’s den’, ‘Loving lady blues’ (1927),
and ‘Jailhouse blues’ (1927) show an incredibly exible approach to 12-bar timing and construction, with
others such as ‘Hesitation blues’ (1927) (based on W. C. Handy’s ‘Hesitating blues’, 1915), ‘I’m still sitting
on top of the world’ (1931) (based on the Mississippi Sheikhs’ 1927 hit ‘Sitting on top of the world’) and
‘Midnight special blues’ (1927) displaying a more orthodox approach to timing and song structure. Of his 22
recordings, 13 display an AAB verse, 12-bar structure. That Collins’s self-composed blues are performed in a
constant state of rhythmical ux, yet pre-written tunes from the Church, string bands, and folk ballad
traditions are performed in precise, metrically formal manners, indicates the importance of elastic timing
to performance in the country blues idiom.

Table 18.6 Comparison of stanza lengths in Charley Pattonʼs ʻPony bluesʼ C18.T6

Song Length of tonic line 1 (in Length of subdominant line 2 Length of dominant line 3 Total no. of
section bars) (in bars) (in bars) bars

Intro 1 N/A N/A 1

Stanza I 4½ 4½ 5 14 ½

Stanza II 3 5 5 13

Stanza III 3 5 5 13

Stanza VI 5 4½ 5 14 ½

Stanza VII 3 5 5 13

Stanza VIII 4 5 4 13

‘Loving lady blues’ is a prime example of Collins’s performance style. He often inserts long slide ‘responses’ C18.P70
to his vocal lines that extend the stanzas beyond the regular 12-bar structure, before righting himself with
the aid of a few stock guitar phrases that act as closure to these improvised passages. These can stretch
verses by up to 16½ bars in length. Like Patton, Collins also shows a marked habit of speeding up during his
performances. ‘Jailhouse blues’ drives from 89 bpm to 116 bpm over its 2 min. 44 sec. duration. It was
perhaps a natural tendency of these performers to increase the tempi of the songs they played for dramatic
15
e ect. This would be e ective in dance scenarios, where the performer desires to increase the intensity of
the dance over the course of a number. However, this acceleration might also be due to nerves on the part of
the performer being in front of a recording machine. The recording environment at this time might be in
warehouses or hotel rooms, and rather alien to performers who were used to playing for live audiences at
dances (Dixon and Godrich 1970). Such sterility might make it hard for performers to feel at ease and nd
their natural tempo, either by rushing or by starting slowly. Otherwise, such acceleration over the course of
p. 382 a commercial recording in the 1920s could be symptomatic of the performer knowing that the brief 3½
minutes of the 78rpm disc required them to rattle through their performance at haste in order for it to be
captured (Katz 2004: 34–35; Philip 2004: 36–38).
Robert Johnson C18.S10

To the wider public, country blues is largely de ned by the gure of Robert Johnson. Born in Hazelhurst, C18.P71
Mississippi, he met an untimely end aged just 27, after recording 29 sides that fell short of commercial
success in his lifetime but posthumously accorded him a fame unrivalled by his contemporaries. He was
recording after the deaths of Je erson and Patton, when country blues was in a commercial decline
following the Depression and the shift to more urban tastes amongst Black consumers. Thus, it is a small

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
miracle that Johnson was recorded at all, and his work has grown in stature since being picked up by artists
such as Elmore James, Led Zeppelin, the Rolling Stones, and Cream. Here, I am less concerned with
Johnson’s time-keeping (which tends to follow a more orthodox approach to metre than other solo blues
artists discussed above) than with postwar listeners’ perceptions of time-keeping and how such prejudices
display a view of country blues as ‘folk art’ despite artists’ attempts to be commercial stars.

One of the interesting side-notes to the legacy of Johnson’s recordings is the continued contention that they C18.P72
have been issued at the wrong speed. 78rpm discs were often recorded at inconsistent speeds, ranging from
60 to 90rpm (Leech-Wilkinson 2009: 50). Some have argued that Johnson’s discs were recorded too fast,
and that by slowing down the recordings such that Johnson sounds a semitone lower in pitch (and of course
delivers the material more slowly), he sounds more natural and authentic (Wilde 2010). There are several
problems with this thesis. Firstly, Johnson’s output was recorded at four sessions over a period of eight
months. The likelihood that the same technical imperfection a ected all of these recordings seems unlikely.
Furthermore, the three available photographs of Johnson show him playing a parlour guitar: being a slightly
shorter-scaled instrument, the slackened strings would su er with tuning problems at low tensions, and
the fact that Johnson has a capo on the third fret of his guitar in one photo corroborates with the pitch that
we hear on his recordings.

The reasons why people have contended that Johnson might have been recorded at the wrong speed are C18.P73
nevertheless fascinating for providing insight into the mentality of revisionist blues histories. They also
pose further questions regarding country blues recordings and the notion of time. Johnson played the guitar
in a fast, bright, attacking style that owes much to the legacy of blues pianists (a further corollary with these
recordings comes from Johnson’s keening croon on numbers such as ‘From four til late’, 1937, and ‘Love in
vain’, 1937, that recall the currently unfashionable style of Leroy Carr and Peetie Wheetstraw). Simply put,
Johnson slowed down ts the perception of a more harrowed, mournful performer than the vivacious
energy and precision of Johnson at standard speed exhibits. The lower-pitched, slower delivery of his
p. 383 performances places him as a kindred performer with such singers as Patton and Son House, whilst it is
evident from Johnson’s recordings that these were only some of his in uences, and that to recast him as
being strictly the progeny of such singers denies him his autonomy as a singular artist.

Johnson was a clear student of recorded blues, basing many of his songs and arrangements on pre-existing C18.P74
songs recorded by the likes of Wheetstraw, Carr, Son House, and Skip James. The speed and precision of his
performances, including the ragtime-in uenced picking style, suggest someone with a knowledge and
a nity with the piano blues recordings that came in the wake of Leroy Carr. Whilst some would rather
believe the mythology of Johnson selling his soul to the devil for the ability to play the guitar (Wald 2004:
265–276), his recorded output (including the near-identical arrangements in many of the available
alternate takes) suggest Johnson to be a craftsman, an artist in uenced profoundly by the blues recordings
that preceded him. He is, in essence, the rst major bluesman whose repertoire is a product of the recording
industry. He rhythmically re ned country blues into more standardized bar durations and regular and
consistent accents, and clearly sculpted his songs, as evidenced by the alternate takes available, in a way
that contrasts with Je erson’s loose revisions of ‘Matchbox blues’. What is interesting is that his
posthumous fame as the archetypical bluesman could have helped further in uence subsequent performers
to adopt his crisp rhythmical regularity, and made him more palatable to White rock musicians. What is also
noteworthy is that Johnson was largely a forgotten gure at the time of his death, his music a commercial
casualty in the face of Black America’s increasingly urban demographic, the rise of swing and R & B, the
electric blues, and the refutation of solo acoustic ‘downhome’ blues in favour of styles stewed in the
avours of the city.

When Columbia Records’ John Hammond tried to track down Johnson to appear at the legendary From C18.P75
spirituals to swing concert at Carnegie Hall in 1939, having been impressed by some of his already obscure
78s, Johnson was already dead. His replacement at the concert was Big Bill Broonzy, a singer with an all-

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
round ability to sing ballads, R & B, folk songs, and blues who became an icon of the 1950s folk revival. But
Johnson was featured at the concert: a lone phonograph was wheeled on stage and Johnson’s music curled
into the hall. Already a ghost, Johnson’s posthumous legend had just begun, a dei cation that led to the
release of compilations of his music, the King of the Delta Blues singers LPs in the 1960s and 70s, his songs
being performed by rock groups such as the Rolling Stones, Canned Heat, and Cream, and nally his
complete recordings being issued in 1990 to substantial media attention (Pearson 1992): the set won a
Grammy Award for ‘Best historical album’ in 1991. Consumers seemingly bought into the ‘devil at the
crossroads’ mythology, and praised Johnson as an atavistic avatar whilst neglecting to acknowledge his
songcraft and the deft concision of his performances, as the primitivist mythology of country blues became
a key perspective from which Johnson, and the whole of country blues, came to be de ned in the popular
consciousness (Wald 2004: 264–276).

p. 384
Conclusions C18.S11

Leroi Jones (2002[1963]) argues that an important relationship has existed between Black music’s C18.P76
development and the historical trajectory of African American social progress. I take my cue from his
assessment of the importance of a historical analysis of Black musical culture in elucidating how people have
come to view the music today. The White relationship with the blues is important in creating a romanticized
obfuscation of Black music culture’s history, which is illuminated through a comprehension of the critical
and public lauding of the male-dominated world of the country blues.

As I have shown, African American country blues recordings exhibit some strikingly exible approaches to C18.P77
timing. The standard 12-bar formulation for blues is treated as a launchpad for extemporizations on the
form, and for uid musical performances that rarely repeat the same verse structures either within a given
performance (evidenced by Charley Patton’s ‘Moon going down’ and ‘Pony blues’) or between separate
versions of the same song (as seen with the example of Blind Lemon Je erson’s ‘Matchbox blues’).
Irregular song-structuring and time-keeping is evident in recordings by several artists, where the
formalized blues structures familiar to millions through the orchestrated recordings, sheet music, and
piano rolls of W. C. Handy’s ‘classic’ blues formulas are recast by country blues artists into new forms of
performance. The recordings that have enshrined these performances have their part to play in shaping
these idiosyncrasies. The time limitations of a 78rpm record are such that we catch a portion of the blues
performance as understood from its more extemporized or elongated forms as found on taped recordings of
blues singers from the 1950s onwards. The recordings themselves become subjects of revisionist debate, as
the metrically precise songcraft of an artist such as Robert Johnson is reimagined at slower tempos that
keep his artistry frozen in an earlier time, rather than representing him as a forward-looking, innovative
musician (Wald 2004). The idiosyncratic timing of the Golden Age of country blues is seen as inherently
authentic, showing these musicians as true folk musicians unhindered by the trappings of commercial
culture—an irony indeed, given that several musicians such as Je erson and Patton were successful
professional performers and recording stars, and that the largest body of recorded country blues from the
era comes not from the ethnographer’s archive but from the commercial output of record labels. Notions of
authenticity in the blues are still highly racialized. It is the province of ‘Blackness’ that provides providence
(Gatchet 2012), and whilst White audiences seem to claim ownership of the country blues by controlling its
marketplace and acting as its principal consumer base, the practice of country blues performance is still
seen as a Black enterprise.

The spell cast is the enchantment of authenticity, and it is aided greatly by the speci c technologies with C18.P78
which the blues was recorded. The narrow sound frequencies reproduced by the early 78rpm shellac records,
the surface noise that is endemic to the format, the compressed sound of a single monaural microphone,
p. 385 and the richness of analog sound recording are part of their appeal. However, these recording technologies

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
ultimately had limitations that have distorted how blues performers could perform for consumers of
78rpm records. Issues regarding timing of the recordings analysed here could be seen to be a ected by these
recording strictures that took performers away from their usual spaces of performance (Katz 2004; Leech-
Wilkinson 2009; Philip 2004). The perceived limitations of the format—coupled with preconceptions
relating to the artists—have enabled the issue of the speed at which Johnson’s recordings were made to
become a topic of debate (Wald 2004). Music can be ‘rei ed and treated as a xed cultural good’ (Cruz 1999:
22), and it is hard to depose these assumptions and received categories once they are established.

Notes
1. The male-dominated domain of acoustic blues music was given the title ʻcountry bluesʼ by Samuel Charters in his 1959 C18.N1
book and LP release of the same name. Whilst Je Todd Titon proposes ʻdownhomeʼ blues (Titon 1977: 317–318), I feel
that neither title is apt, as both reinforce a revisionist distinction between the male acoustic blues and the vaudeville-jazz
of the ʻclassicʼ blues singers of the 1930s. Female stars such as Bessie Smith and Victoria Spivey paved the way for the
recording of the male acoustic blues and inspired much of the latterʼs repertoire. Perhaps without realizing it, these
writers are still imposing categories upon the blues that are not of the genre. I use such terms as ʻcountryʼ and
ʻdownhomeʼ blues with this in mind, recognizing the inability of scholars to divorce their discussions from the popular
history of the blues which has restricted their vocabulary since the resurgence of interest in rural blues began in the 1950s.

2. For more on blues as a niche music in black America, see Nicholson (1998) and Gatchet (2012). C18.N2

3. The reworking of Skip Jamesʼs ʻDevil got my womanʼ (1931) into Robert Johnsonʼs ʻHellhound on my trailʼ (1937) is C18.N3
instructive of this process. See Wald (2004: 150–151, 171–172).

4. Acoustic recording was non-electrified, and required singers and musicians to perform into wax horns that would etch C18.N4
their performances into wax. Electric recording allowed microphones to amplify the signal etched into the wax, thus
allowing quieter and more nuanced performances to be captured e ectively (Leech-Wilkinson 2009: 8).

5. Lomax (1934; 1947: 115–140). Recordings are available at the Alan Lomax Archive: http://research.culturalequity.org/get- C18.N5
audio-ix.do?ix=sessionandid=PR47andidType=abbrevandsortBy=abc (accessed 3 Apr. 2019).

6. See Fat Possum Records (https://www.fatpossum.com/) for a survey of Mississippi blues in the new millennium; also C18.N6
Nicholson (1998).

7. As mentioned above, W. C. Handyʼs popularizing of the blues in the 1910s spread the form across America. White folk C18.N7
artists such as Dock Boggs, Jimmie Rodgers, and Charlie Poole all recorded songs in the blues idiom during the 1920s.

8. The Mississippi Delta of blues lore consists of the inland delta between the Yazoo and Mississippi rivers. C18.N8

9. Examples include Charley Pattonʼs ʻHigh water everywhereʼ (1930), Furry Lewisʼs ʻKassie Jonesʼ (1928), and Son Houseʼs C18.N9
ʻMy Black Mamaʼ (1930).

10. Version 1 (OKeh 8455, 14 Mar. 1927), version 2 (Paramount master no. 4424-2, Apr. 1927), and version 3 (Paramount master C18.N10
no. 4424-4, Apr. 1927).

p. 386 11. ʻFloating versesʼ are stock lyrical verses that appear in many di erent blues performances, sometimes with minor C18.N11
alterations but o en copied wholesale (Oliver 1960). Examples of country blues verses that originated in the classic blues
repertoire include ʻGood morning blues, blues how do you do / Iʼm doing alright, good morning, how are youʼ—a lyric that
bridges from Bessie Smith to Leadbelly (Sackheim 1969: 52).
12. Yazooʼs release came a er the Origin Jazz Libraryʼs releases The Immortal Charlie [sic] Patton Number 1 (1962) and The C18.N12
Immortal Charlie [sic] Patton Number 2 (1964).

13. The ʻJinx/Maggie Campbellʼ pattern is an open-G tuning consisting of a descending 4-note bass run o en using octaves on C18.N13
the 4th and 6th strings (Evans 1982: 146–149, 245).

14. The song was also the first to be issued by Blind Lemon Je erson for Paramount Records, under the pseudonym ʻDeacon C18.N14
L. J. Batesʼ, in 1926.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
15. This tendency is particularly evident in Pattonʼs recording of ʻI shall not be movedʼ (1929), which starts at 77 bpm before C18.N15
resolving at a tempo of 118 bpm. ʻPony bluesʼ similarly starts at a stately 73 bpm but ends up galloping along at 104 bpm.
References C18.S12

Appadurai, A. (1988). Place and voice in anthropological theory. Cultural Anthropology 3(1): 16–20. C18.P79
Google Scholar WorldCat

Aubert, L. (2007). The music of the other: New challenges for ethnomusicology in the global age. Ashgate. C18.P80
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
Charters, S. (1959). The country blues. Rinehart. C18.P81
Google Scholar Google Preview WorldCat COPAC

Cook, I., and Crang: (1996). The world on a plate: Culinary culture, displacement and geographical knowledges. Journal of C18.P82
Material Culture 1(2): 131–153.
Google Scholar WorldCat

Cruz, J. (1999). Culture on the margins: The Black American spiritual and the rise of American cultural interpretation. Princeton C18.P83
University Press.
Google Scholar Google Preview WorldCat COPAC

Dahl, B. (2016). The art of the blues: A visual treasury of Black musicʼs Golden Age. University of Chicago Press. C18.P84
Google Scholar Google Preview WorldCat COPAC

Davis, A. (1998). Blues legacies and Black feminism: Gertrude ʻMaʼ Rainey, Bessie Smith, and Billie Holiday. Pantheon. C18.P85
Google Scholar Google Preview WorldCat COPAC

Dixon, R. M. W., and Godrich, J. (1970). Recording the blues. November Books. C18.P86
Google Scholar Google Preview WorldCat COPAC

Dougan, J. (2006). Objects of desire: Canon formation and blues record collecting. Journal of Popular Music 18(1): 40–65. C18.P87
Google Scholar WorldCat

Evans, D. (1982). Big road blues: Tradition and creativity in the folk blues. University of California Press. C18.P88
Google Scholar Google Preview WorldCat COPAC

Fabian, J. (2002). Time and the other: How anthropology makes its object. Columbia University Press. C18.P89
Google Scholar Google Preview WorldCat COPAC

Fahey, J. (1970). Charley Patton. November Books. C18.P90


Google Scholar Google Preview WorldCat COPAC

Gatchet, R. (2012). ʻIʼve got some antique in meʼ: The discourse of authenticity and identity in the African American blues C18.P91
community in Austin, Texas. Oral History Review 39(2): 207–229.
Google Scholar WorldCat

Gioia, T. (2008). Delta blues: The life and times of the Mississippi masters who revolutionized American music. Norton. C18.P92
Google Scholar Google Preview WorldCat COPAC

Hamilton, M. (2007). In search of the blues: Black voices, White visions. Jonathan Cape. C18.P93
Google Scholar Google Preview WorldCat COPAC

Jones, L. (né Baraka, A.) (2002[1963]). Blues people: Negro music in White America. Perennial. C18.P94
Google Scholar Google Preview WorldCat COPAC

p. 387 Katz, M. (2004). Capturing sound: How technology has changed music. University of California Press. C18.P95
Google Scholar Google Preview WorldCat COPAC

Keil, C. (1966). Urban blues. University of Chicago Press. C18.P96


Google Scholar Google Preview WorldCat COPAC

Kuhn, J. (2005). Audiotopia: Music, race and America. University of California Press. C18.P97
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
Leech-Wilkinson, D. (2009). The changing sound of music: Approaches to studying recorded musical performances. CHARM. C18.P98
Google Scholar Google Preview WorldCat COPAC

Locke, A. (1939). The Negro and his music. Arno Press. C18.P99
Google Scholar Google Preview WorldCat COPAC

Lomax, A. (1934). ʻSinfulʼ songs of the southern negro: experiences collecting secular folk-music. Southwest Review 19(2): 105– C18.P100
131.
Google Scholar WorldCat

Lomax, J. (2017[1947]). Adventures of a ballad hunter. University of Texas Press. C18.P101


Google Scholar Google Preview WorldCat COPAC

Lornell, K. (2000). Blind Lemon meets Leadbelly. Black Music Research Journal 20(1): 23–33. C18.P102
Google Scholar WorldCat

Lott, E. (1993). Love and the : Blackface minstrelsy and the American working class. Oxford University Press. C18.P103
Google Scholar Google Preview WorldCat COPAC

McGinley: (2014). Staging the Blues: From tent shows to tourism. Duke University Press. C18.P104
Google Scholar Google Preview WorldCat COPAC

Myers, F. (1995). Representing culture: The production of discourse(s) for Aboriginal acrylic paintings. Cultural Anthropology 6(1): C18.P105
26–62.
Google Scholar WorldCat

Narváez: (2002). The influences of Hispanic music cultures on African-American blues musicians. Black Music Research Journal C18.P106
22 (supplement: Best of BMRJ): 175–196.
Google Scholar WorldCat

Nicholson, R. (1998). Mississippi: The blues today! Blandford. C18.P107


Google Scholar Google Preview WorldCat COPAC

Oliver: (1960). Blues fell this morning: Meaning in the blues. Cambridge University Press. C18.P108
Google Scholar Google Preview WorldCat COPAC

Oliver: (1984). Songsters and saints: Vocal traditions on race records. Cambridge University Press. C18.P109
Google Scholar Google Preview WorldCat COPAC

Ottenheimer, H. J. (1979). Catharsis, communication and evocation: Alternative views of the sociopsychological functions of C18.P110
blues singing. Ethnomusicology 23(1): 75–86.
Google Scholar WorldCat

Patton, C. (1971). Founder of the Delta blues. Yazoo Records. C18.P111


Google Scholar Google Preview WorldCat COPAC
Pearson, B. (1992). Standing at the crossroads between vinyl and compact discs: Reissue blues recordings in the 1990s. Journal C18.P112
of American Folklore 105(416): 215–226.
Google Scholar WorldCat

Peterson, R. (1997). Creating country music: Fabricating authenticity. University of Chicago Press. C18.P113
Google Scholar Google Preview WorldCat COPAC

Petrusich, A. (2014). Do not sell at any price: The wild, obsessive hunt for the worldʼs rarest 78rpm records. Scribner. C18.P114

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Philip, R. (2004). Performing music in the age of recording. Cambridge University Press. C18.P115
Google Scholar Google Preview WorldCat COPAC

Ramsey, G. P. (2003). Race music: Black cultures from bebop to hip-hop. University of California Press. C18.P116
Google Scholar Google Preview WorldCat COPAC

Sackheim, E. (1969). The blues line: Blues lyrics from Leadbelly to Muddy Waters. Thunderʼs Mouth Press. C18.P117
Google Scholar Google Preview WorldCat COPAC

Skinner, K. (2006). ʻMust be born againʼ: Resurrecting the anthology of American folk music. Popular Music 25(1): 57–75. C18.P118
Google Scholar WorldCat

Szwed, J. (2011). The man who recorded the world: A biography of Alan Lomax. Arrow Books. C18.P119
Google Scholar Google Preview WorldCat COPAC

Titon, J. T. (1977). Thematic pattern in downhome blues lyrics: The evidence on commercial phonograph records since World C18.P120
War II. Journal of American Folklore 90(357): 316–330.
Google Scholar WorldCat

Titon, J. T. (1994) Early down home blues (2nd edn). University of North Carolina Press. C18.P121
Google Scholar Google Preview WorldCat COPAC

Various artists (1959). The country blues. RBF Records. C18.P122


Google Scholar Google Preview WorldCat COPAC

Various artists (1962). Really! The country blues 1927–1933. Origin Jazz Library. C18.P123
Google Scholar Google Preview WorldCat COPAC

Various artists (2001). The great women blues singers. Recording Arts. C18.P124
Google Scholar Google Preview WorldCat COPAC

Various artists (2013). Parchman Farm: Alan Lomaxʼs photographs and field recordings, 1947–1959. Dust-to-Digital. C18.P125
Google Scholar Google Preview WorldCat COPAC

p. 388 Wald, E. (2004). Escaping the Delta: Robert Johnson and the invention of the blues. Amistad. C18.P126
Google Scholar Google Preview WorldCat COPAC

Wardlow, G. D. (1998). Chasinʼ that devil music: Searching for the blues. Backbeat Books. C18.P127
Google Scholar Google Preview WorldCat COPAC

Weissman, D. (2006). Which side are you on? An inside history of the folk music revival in America. Continuum. C18.P128
Google Scholar Google Preview WorldCat COPAC

Wilde, J. (2010). Robert Johnson revelation tells us to put the brakes on the blues. Guardian. C18.P129
Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470592 by National Science & Technology Library user on 26 May 2023
<https://www.theguardian.com/music/musicblog/2010/may/27/robert-johnson-blues>.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
CHAPTER

19 Metrical Displacement and Group Interaction


C19 in

‘Evidence’ by the Thelonious Monk Quartet 


Ryan D. W. Bruce

https://doi.org/10.1093/oxfordhb/9780190947279.013.19 Pages 389–C19.P89


Published: 08 December 2021

Abstract
Jazz pianist Thelonious Monk is known for his rhythmically complex compositions and
improvisations. His typical 32-bar AABA form pieces provide a framework of musical norms in terms
of harmonic movement, and thus a point of reference for the harmonic rhythm to be displaced.
‘Evidence’ is exemplary of Monk’s displaced rhythms, which creates a sense of metrical shifts during
the head arrangement. Composed c.1948 and a frequent piece of Monk’s performance repertoire into
the 1970s, a transcription and analysis of the recording from Thelonious Monk Quartet Plus Two at the
Blackhawk demonstrates a con icting sense of metre between band members of the quartet. This
chapter investigates how the musicians negotiated metrical discrepancies in the composed section and
the saxophone solo in terms of group interaction. A mediation of time is demonstrated by a displaced
metre in the drums and each musician’s performative response to the discrepancy by providing
musical signals during the course of performance. Through analysis of this performance, the chapter
examines the way in which the musicians arbitrated a decisive point of reference within a confounding
performance of overturning the beat. This chapter contributes to understanding rhythm and metre
through improvisatory processes, augmenting scholarship on jazz through its analysis of the temporal
constituents of group interaction.

Keywords: Thelonious Monk, group interaction, jazz analysis, metrical displacement, metrical dissonance,
rhythmic displacement
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

A live performance of ‘Evidence’ by Thelonious Monk’s small ensemble in 1960 proved to test their C19.P1
communicative abilities as a group (Monk 1987). Captured on Monk’s Thelonious Monk Quartet Plus Two at
the Blackhawk recording, this rhythmically obscure composition causes the drummer, Billy Higgins, to turn
the beat around at the end of the head arrangement. The ensuing interplay between performers
demonstrates di erent readings of the metre—the musicians continually change between the original beat
emphasis and another where the metre is out of phase. Enculturated listeners can usually hear the metrical
dissonance between the parts, and the process of metrical displacement as the metre seems to shift back
and forth by a quarter-note pulse. This occurs eight times in the head and the tenor saxophone solo. It takes
the group a full minute to negotiate their di erences, communicating in di erent ways to arbitrate a
common beat emphasis and carry the performance forward. This chapter investigates the head arrangement
and the saxophone solo on the recording, with a focus on the performers’ rhythmic interaction as it relates
to metre. Following a discussion of issues relating to rhythmic gestures in jazz performance, the chapter

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
provides an analysis of the di erent musical signals deployed by the musicians in this performance, and the
role of these signals to achieve metrical coherence during the improvisation.

1
Analyses of rhythm and metrical shift in jazz research have tended to focus on individual performers. Some C19.P2
analysts suggest that changes in metre may be heard according to di erent metric groupings such as threes
2
or ves (and, in some instances, half-beat divisions). For Monk’s music, Mark Haywood provides an
additional stave below the piano part, in which rhythmic groupings are labelled as separate and isolated
p. 390 metric groupings (1995). Although these groupings may have been intended in pre-planned
composition, I hesitate to impose these kinds of metric readings on the saxophone solo played in this
performance. With a 4/4 swing groove as the basis for a collective entrainment, negotiation at the time of
performance is better understood by the metre’s equal division into two beats or divided once again for
adjustment at the beat level. In other words, temporal adjustments may simply be one or two beats in
duration, rather than rhythmic gestures that imply longer odd-numbered metrical groupings.

Stefan Love (2013) analyses metrical shifts according to displacements of the beat. Although it is beyond the C19.P3
scope of this chapter, his writing includes a summary of theory and method from Krebs (1999) and Mirka
3
(2009) that are apt for jazz analysis. I similarly adopt Krebs’s descriptions of displacement dissonance (a
con ict arising from the o set of two layers—herein instrumental parts—of divisible cardinality), which
are subcategorized as direct dissonance (times of hearing the con ict of o set layers) and indirect
dissonance (the brief moment when a listener perceives a metrical shift) (1999: 31–48). I also support the
idea that subliminal dissonances—when musical features express a beat emphasis that is displaced from the
original metre (Krebs 1999: 46)—may result in moments of local metrical consonances, experienced as
such by listeners and committed by at least one band member at the time of performance (Love, 2013).

The rhythmic complexities in improvised jazz performance are not however solely due to metrical shifts and C19.P4
temporal displacements of various kinds but also down to the considerable uidity of timing that adds ‘feel’
to the performance. In ‘Evidence,’ the saxophone’s changing time-feel comes from delayed notes and
4
playing on top or behind the beat. Such microtimings have been studied in various ways from the
participatory discrepancies that Charles Keil describes (1995) to more psychological investigations of
expressive timing (see Collier and Collier 2002). Vijay Iyer writes about the production and perception of
microtiming in jazz, by discussing the expressive qualities in a recording of ‘I’m Confessin’’ by Monk:

The question of whether Monk ‘intended’ to play this in exactly this way is a pejorative one, akin to C19.P5
reifying the role of ‘mistakes’ in jazz […] It is never clear what is ‘supposed’ to happen in
improvised music, so it makes little sense to talk about mistakes. From the perspective of an
improvisor, the notion of a mistake is supplanted by the concept of interaction with the structure
suggested by the sonic, physical, and temporal environment.

(Iyer 2002: 408)

The topic of mistakes in Monk’s music pervades the literature about him, and recent analyses treat them as C19.P6
a desired outcome of his conception. A focus on rhythm was a primary concern for Monk: rhythmically
displaced phrases with economical use of simple ideas and silence are one of his trademarks. With respect to
the metre, Monk may change emphasis to the backbeat, or the rhythm section may adjust to a rhythmically
displaced idea creating a perceived shift of the metre. When this process transpires, the rhythms ‘reach a
point of complexity that challenges the listener’s ability to hold on to the meter’, and in turn, the ‘meter is
overturned in the face of confusing and contradictory signals’ (Feurzeig 1997: 2).

p. 391 Since the 1990s, the recognition that displacements, whether mistakes or intentional shifts, are not simply C19.P7
the preserve of individual players but a matter of group dynamics has led analysts to pay more attention to
interaction in jazz performance. Paul Berliner uses the metaphor of conversation to describe improvisatory

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
processes for group cohesion (1994). He explains that unforeseen events (e.g. deviations from harmony,
metre, or form) require strategies for the musicians to nd a common ground (1994: 379–383). Monk’s
music is known for its complex rhythms in this context, explained by drummer Leroy Williams:

Some players can stretch the time to that ne line of almost turning the beat around, but they can C19.P8
always come back. For example, with [bassist] Wilbur Ware in Monk’s band, they would play so
close to that thin line rhythmically that, if you weren’t careful, you’d nd yourself playing on ‘one’
and ‘three,’ instead of ‘two’ and ‘four.’ If you weren’t careful, you’d be right o it.

(Berliner 1994: 381)

Band members nd musical ‘saves’ in such occurrences—attempts to get the performance back into C19.P9
alignment—such as providing formal cues to each other, continuing with the harmonic form until the
erroneous musician falls into place, or collectively adjusting to the particular error (Berliner 1994: 382).
Beyond Berliner’s work, Ingrid Monson (1996) has also analysed interaction in jazz and writes extensively
about the metaphors of conversation and music-as-language, pointing to the need for mutual support in
5
performance when things go awry. Related to timing, she demonstrates how musicians may become out of
phase with each other requiring interactive adjustments to the metre and form (1996: 152–71). Robert
Hodson similarly nds an occurrence in Thelonious Monk’s solo of ‘Misterioso’ (1958) when the musicians
are implying harmonic rhythm on di erent beats and add an extra bar into the form to become in sync
become in sync (2007: 91–95).

Metrical displacement and moments of local consonance may both be a result of entrainment between C19.P10
players. Do man (2013) explains how the groove in jazz is a shared coherence of time for bandmates during
improvisation. The groove is not simply a basis for interaction but ‘considerably expressive’: deviations
from the ‘norm’ represented on a score require a more uid understanding, and ‘perhaps the notion of
equilibrium within an entrained relationship can stand as a norm’ (Do man 2013: 78, emphasis original).

My analysis of ‘Evidence’ shows how the musicians moved in and out of the shared coherence of a swing C19.P11
groove. The overturned metre is a consequence of musical signals performed by the musicians and then
interpreted (or misinterpreted) by each other. The salient signals have been examined from a listener’s
perspective—in other words, my recurrent listenings to the piece undertaken to clarify how these metrical
discrepancies occurred and were solved through musical action. Metrical shifts such as those in this piece
are usually ambiguous and can be read in di erent ways (analogous to the duck/rabbit illusion!). In this
case, the shifting metre between the players revolves around Higgins’s initial error in turning the beat
round (playing beat two as beat one of the bar) during the head of the tune. This initial error then leads to
considerable ambiguity, and the performance can be heard principally both in terms of what I call the
p. 392 ‘shifted paradigm’ (the metrical ‘good sense’ imposed by the listener and/or the other players) vis-à-vis
the ‘original paradigm’ (what was actually played in reference to the rst beat of the tune; the ‘turned-
6
around’ beat). The musical signals used by the players suggest that the musicians used a limited number of
solutions in attempting to come back together when a sense of the metre was in disagreement. Clearly, this
analysis cannot identify the intentions of the musicians in the moment, but it does o er a highly plausible
account of how the musicians recovered their temporal equilibrium.
Analysis of ʻEvidenceʼ C19.S1

‘Evidence’ is a medium swing, 32-bar, AABA song-form piece in E♭ major, based on the chord progression C19.P12
of the jazz standard ‘Just you, just me’. Monk rst recorded it in 1948, and it was a consistent part of his
performance repertoire thereafter. The Blackhawk recording (Monk 1987) is played at 190 bpm, and has the
overall form of an A-section solo piano introduction followed by the standard pattern of head–solos–head,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
with a two-chorus tenor saxophone solo as the rst. The group includes three regular members on tour
from New York: Monk on piano, Charlie Rouse on tenor saxophone, and John Ore on bass. Billy Higgins was
the local San Francisco hire on drums (although an internationally celebrated player).

7
The saxophone, piano, bass, and drums in the head and the saxophone solo are the object of the analysis. C19.P13
Eighth-note pairs in the gures represent swung rhythms, or long-short durations somewhere around the
rst and third onsets of a triplet gure. The drums provide a standard swing pattern on the ride cymbal: a
8
quarter-note followed by two swing eighth notes is traditional for most bop recordings. Higgins (on
drums) turns the beat around, creating a displaced swing pattern in the rst chorus of the saxophone solo.
The transition from the head to the solo is a starting point to investigate the displaced metre, as shown at
bar 41 in Figure 19.1. The top stave shows the original paradigm, which was likely interpreted as the shifted
paradigm on the bottom stave: a standard swing pattern on the ride cymbal.

Figure 19.1

Example of the shi ed and original paradigm for the drum part. C19.F1

p. 393 In the original paradigm on the top stave, the swing pattern is displaced with beat 2 sounding as the C19.P14
downbeat, shown in the shifted paradigm on the bottom stave and annotated with an arrow. Experienced
listeners may expect to hear a hi-hat on the rst of each eighth-note pair. Whether due to the performance
or the recording quality, the hi-hat is usually absent in this recording and has been excluded from
transcription. Its absence is insigni cant to the swing pattern provided here. The single snare-drum hit is
simply included because it was part of Higgins’s drumming at the time.

We hear Figure 19.1 as a standard swing pattern (i.e. the shifted paradigm on the bottom stave), and its C19.P15
displacement (i.e. the original paradigm) is the ‘mistake’ for the other musicians to respond. Higgins does
not change his drum pattern through the piece, and the other musicians adjust to be in sync later in the solo.
That process is the subject of analysis having looked at the beginning of the piece.

The drums turn the beat around in the A sections of the head arrangement. The solo piano introduction is C19.P16
not metric. It infers a rhythmic quarter-note tactus but it is not con rmed in any metric reading (i.e.
repeated groups of 2 or 3). The group enters to play the head from at bars 9–40. Figure 19.2 is the second A
section and representative of the others. Labels for the staves are as follows: p. (piano), b. (bass), and d.
(drums).

The rst E♭ note on piano is relatively inaudible: the next note, D is heard as the rst note of the melody, C19.P17
and is perceived as the rst note of the section. For the most part, the melody in the A sections is played by
all ensemble members in homorhythm (the saxophone plays the top notes of the piano part with an
ornamental turn in bar 22). The drums articulate the rhythm on the bass drum, with occasional accents on

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
the ride cymbal and snare drum, and lling the space between with reference to a swing pattern of a
quarter-note followed by two swing eighth notes which are interrupted or delayed by the next melody note.
Metre is not provided at the beginning, only rhythmic groupings of three beats, or with a quarter-beat
diminution (labelled 3? in bar 18) or quarter-beat augmentation (bars 19 and 20).

A reference to the pulse is provided in bar 20 with two adjacent quarter-notes; in bar 22 we hear an C19.P18
important structural moment with traditional harmonic rhythm on beats 1 and 3. The subsequent melody
note in bar 23 appears on beat 2 and begins a 4-beat pattern. The drums follow slightly behind the melody,
treating the notes of G# and B in the piano as an anticipation of the downbeat rather than appearing on the
‘and’ of beat 1 as shown in the score. The rhythmic treatment begins a displaced swing pattern in 4/4 time:
waiting for another 4 beats (labelled as 4?), the drums continue with a 4-beat pattern in the B section.

The B section is an example of metrical shift and a local consonance. The melody in the B section—the C19.P19
ascending stepwise movement on the ‘and’ of beat 1—aids in a collective shift of the metre due to the drums
and the bass as shown in Figure 19.3. A reader will not hear the bracketed pairs sounding simultaneously;
the reading requires one to see the original paradigm on each top stave but hear the bottom staves of each
pair.

The displaced swing pattern in the drums suggests that the ‘and’ of beat 1 has become the ‘and’ of beat 4, an C19.P20
anticipation of the next bar’s downbeat. The shifted metre is con rmed in the bass, at times providing
p. 394 chromatic root movement to articulate chord changes in the melody now perceived as anticipations of
beat 1. In the transition to the next A section, the melody note D is heard as a downbeat. The arrows pointing
to the top stave only suggest that the tune has returned to its original paradigm for the last A section
(including Higgins’s three-beat drum patterns from Figure 19.2).
Figure 19.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
The A section and beginning of B section of ʻEvidenceʼ; shows the head arrangement with drum patterns played in 3 or 4 beats. C19.F2

The way that Higgins previously ‘turned the beat around’ is repeated in this last A of the head. The displaced C19.P21
metre is prepared two bars before the solo. In Figure 19.4, the rst E♭ on the piano is treated as a downbeat
in bar 39 (the shifted paradigm). The melody lends itself to the shifted metre: the nal melody note in bar 39
p. 395 sounds as an anticipation of the downbeat, and the drums continue with a seamless yet displaced swing
p. 396

pattern. The piano comping in bar 41 is shown to ‘force’ our perception of the metre back to its original
paradigm, sounding as a direct metrical dissonance with the drums.
Figure 19.3

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
The melody of the B section (original paradigm on the top stave) has been shi ed, showing where rhythms may be perceived as C19.F3
anticipations of downbeats (the shi ed paradigm on the bottom stave).
Figure 19.4

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Displaced swing pattern at the end of the head. C19.F4

The following analysis looks at the instruments separately to indicate the signals each musician provides on C19.P22
the bandstand. The results are then combined to see how the musicians interacted with those signals to nd
a decisive beat emphasis in accordance with a standard swing groove.

Monkʼs Signals on Piano C19.S2

The saxophone solo begins at bar 41. Aware that the beat was overturned in the drums, Monk introduces one C19.P23
solution by clearly marking the downbeat of each bar with whole-note chords for ve bars. Figures 19.5 and
19.6 compare the melody from the head on the top stave and the piano comping on the bottom stave.
Comping similar to the melody is indicated with enclosures. The comping references the strong harmonic
rhythm from the melody in bar 46 (analogous to bar 6 of the A section). The last bars of the A section clearly
state strong beats 1 and 3, with an anticipation of beat 1 for the last bar, as is conventional. Typically of
Monk’s comping strategies, he plays the melodic rhythm almost verbatim in the following A section from
9
p. 397 bar 49 to 56. The rhythms displaced by a beat are indicated with arrows in Figure 19.5. Monk’s signal to
the other band members, however, is a clear restatement of the melody in the last half of the second A
section.
Figure 19.5

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Comparison of the melody and Monkʼs piano comping, bars 41–56, indicated by the enclosures. C19.F5

The transition to the B section has a 2-against-3 polyrhythm that leads into the melody, as shown in Figure C19.P24
19.6. Monk’s signal may be understood in context of a restated melody; as I demonstrate later, the black
chords on the ‘and’ of beat 1 aid in turning the beat around much like the B section in the head. Monk’s
comping in the last A section of the chorus (bars 65–72) is also a restatement of the melody.

The rst A section in the next chorus is shown in Figure 19.7. Monk continues to reference the melody, again C19.P25
providing a clear statement metre in bar 78 (analogous to bar 6 with chord changes on beats 1 and 3). He
clearly articulates anticipations of strong beats in the II–♭II–I progression (in E♭) to conclude the section,
and adds a beat into the form after bar 79 (shown as a 1/4 bar).
p. 398 Figure 19.6

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Comparison of the melody and Monkʼs piano comping, bars 57–72, indicated by the enclosures. C19.F6

Higginsʼs Signals on Drums C19.S3

Higgins maintains a displaced swing groove on drums. His musical signals are primarily from the snare C19.P26
drum. Accounting for the displacement, I analysed the occurrences of snare-drum hits in the shifted
paradigm, for example, treating the snare on the ‘and’ of beat 4 in Figure 19.8 rather than the original
metric paradigm in the top stave.

In the 32 bars of the rst solo chorus, there is far more emphasis on strong beats with the snare drum on the C19.P27
‘and’ of 4 or the ‘and’ of 2, traditional rhythmic gestures that anticipate strong beats 1 and 3 respectively.
Table 19.1 shows 15 occurrences of the snare drum on the ‘and’ of beat 4, and 13 on the ‘and’ of beat 2 in
p. 399 comparison to the remaining beats.
Figure 19.7

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Comparison of the melody and Monkʼs piano comping, bars 73–81, indicated by the enclosures. C19.F7

Figure 19.8

Displaced swing groove and its shi ed paradigm with snare drum on the ʻandʼ of beat 4. C19.F8

Table 19.1 Occurrences of the snare drum in the first solo chorus, bars 41–73 C19.T1

Beat Occurrence

1 1

ʻAndʼ of 1 3

2 7

ʻAndʼ of 2 13

3 0

ʻAndʼ of 3 4

4 1

ʻAndʼ of 4 15

p. 400 When a comping pattern includes two or more snare-drum hits per bar, almost all patterns include the C19.P28
‘and’ of 4. That is, the other hits are rarely heard in isolation. Figure 19.9 shows three primary comping
units when grouping the instances of the ‘and’ of beat 4 with the remaining beats. The third unit excludes
stems to represent any combination of those snare drum occurrences.
Figure 19.9

Snare drum comping units in the first solo chorus, bars 41–73, appearing on: (1) the ʻandʼ of beat 4; (2) the ʻandʼ of beats 2 and 4; C19.F9
and (3) the ʻandʼ of beat 4 and any other beat in the bar.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Considering these units to represent articulations of the ‘and’ of beat 4, Table 19.2 subsumes the remaining C19.P29
beats under the ‘and’ of beat 4 according to the instances of the comping units in Figure 19.9. The reduction
yields a swing groove with primary comping patterns anticipating beats 1 and 3.

Table 19.2 Reduction of snare drum comping patterns in the first solo chorus, bars 41–73 C19.T2

Beat Occurrence

1 0

ʻAndʼ of 1 1

2 1

ʻAndʼ of 2 9

3 0

ʻAndʼ of 3 1

4 0

ʻAndʼ of 4 15

The last signal Higgins provides is with the bass drum. It is heard twice in the rst chorus of the solo. This C19.P30
use of the bass drum as an interjection is referred to as ‘dropping bombs’; the rst instance comes on beat 4
in the bar before the B section and is relatively insigni cant. The second time is after bar 72, shown in
Figure 19.10. The transition to the second solo chorus has two bass drum strikes, and a snare drum on beat 1
to emphasize the downbeat of the swing groove. It con rms the metric adjustment in the bass after the
beginning of bar 71, and a possible adjustment on saxophone after bar 72.

It is important to note that the drums did not insert a beat as shown by the 1/4 bar. Higgins did not change C19.P31
p. 401 the swing pattern; rather, it is at this point in the form that the bass and saxophone parts can be seen to
adjust to the drums. The 1/4 bar in the drum part simply means that the beat was being permanently
overturned, and heard as a standard swing groove with the rest of the ensemble.
Figure 19.10

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Bass drum emphasizing the downbeat at the beginning of chorus 2. C19.F10

Rouseʼs Formulas as Signals on Saxophone C19.S4

Formulaic analysis is provided to make sense of Charlie Rouse’s metric emphasis on saxophone. The notion C19.P32
of formulas was rst de ned by Albert Lord (1960) to describe oral transmission of epic poetry, and
subsequently adopted by musicologists to analyse similar repeated motives in oral musical traditions. In
jazz, formulas occur at analogous points in the cyclic form of pieces (such as a 12-bar blues, or an AABA
form piece like ‘Evidence’), and across a repertoire of pieces (Gushee 1991). I expand this concept to include
similar phrases that imply a harmonic function (such as V–I movement) or beat emphasis of comparable
like and kind. The formulas presented here have been aurally veri ed in other recordings of ‘Evidence’ by
10
Rouse.

Formulas are bene cial for analysis because of their exibility. They are not bound by speci c pitches but C19.P33
arise as a group of motives based on similarity. That is, formulas may have slightly di erent pitch content,
or be placed on di erent beats for a rhythmic displacement of phrases.

p. 402 In the Blackhawk recording, Rouse’s formulas occur at analogous times in the A and B sections, as well as C19.P34
times with V–I chord changes. The formulas provide one of two metric signals that may be interpreted by
the rest of the band: (1) an indication of harmonic rhythm by ‘playing the changes’ or (2) rhythmic
displacements at structural moments of the form. The formulas identi ed here are not exhaustive of
Rouse’s playing, and are chosen as representatives that contribute to metrical shifts.

The formula labelled α is typical for V–I chord changes in jazz, and usually resolves to a chord tone. In C19.P35
Figure 19.11, the core of the formula shows the progression of alternating eighth notes in a stepwise descent
11
of the top notes to the resolution. The core is an example showing scale degrees 5–2–4–2 and a resolution
to 3 in E♭ major across the bar line. The harmonic rhythm is implied and a perception of the metre
established through the formula’s cadential emphasis of a strong beat.
Figure 19.11

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Formula α on saxophone characterized by alternating notes that descend and resolve on a chord tone. C19.F11

The core is indicated with brackets for each occurrence. Rouse rhythmically displaces the phrase throughout C19.P36
the recording, such as resolving on weak beats (beat 2 in bars 53 and 69, and beat 4 in bars 79 and 86) or
p. 403 strong beats (beat 3 in bar 75, and beat 1 in bar 94). The rst two instances of the formula appear at
analogous times in the cyclical form, from bars 4–5 of the A section. In bars 52–53, the phrase may be seen
as beginning in B♭7, a tritone substitution of the E7 harmony heard in the piano and bass parts. The
remaining formulas are similar in melodic shape. The four examples from bars 52–80 include root
movement of a descending fth, whereas the example in bars 85–86 includes a resolution to a prolonged
E♭7 harmony. In bars 93–94, the core seemingly resolves to a chord tone; however, the phrase sounds
independent of the D♭7 harmony with its continuation to beat 3 in bar 94. For this analysis, the speci c pitch
content is less important than the cadence point, which serves as a signal for a possible metric reading.

Stating that Rouse was making mistakes with resolutions to weak beats would be incorrect, imposing an C19.P37
assumption that the gure ought to resolve to a strong beat. From the rudimentary de nition of the core,
the notation may incorrectly suggest that the beat is overturned or that a sense of metrical shift occurs with
weak-beat resolutions. Such is not always the case; rather, we can think of the formula as a phrase that
contains a resolution at di erent times in the metric grid. The resolutions to weak beats may simply be
rhythmically displaced phrases, which Rouse would perform often. What can be gathered from this formula
is a set of signals that (1) to the listener, may sound rhythmically displaced or create a perceived metrical
shift, and (2) may have been interpreted by the band, as listeners and performers, to negotiate a sense of
metric continuity.
Figure 19.12

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Formula β on saxophone; the core melodic cell of a descending interval. C19.F12

The second formula labeled β in Figure 19.12 is a melodic cell. The descending interval, usually a minor third C19.P38
as shown in the core, does not have signi cant harmonic function. The core is written rhythmically
according to its rst appearance. Transformations of the core appear as expanded descending intervals in
p. 404 bar 57, beat 3 of bar 60, and through intervallic and rhythmic augmentation in beats 2–3 in bar 92 (each
resolving on the tonic of E♭). The examples from bars 58–60 and 90–92 are aligned to show the similar
pitch content, and that they occur at analogous times of bars 3–4 of the B section.

In bar 45 Rouse is playing on top of the beat; beat 3 is emphasized due to the accent on the rst note, and C19.P39
with microtiming of straight eighth notes. Similar to formula α, the remaining cells occur on di erent beats,
pointing to a uidity of rhythmic gesture within the bar. What makes them formulaic is the rests before and
after the cells, and the similarities between the last four examples: (1) each cell is followed by at least three
beats of rest, and (2) the cells are heard as articulating beat 3 (in a metrical shift), requiring us to reread the
rhythms in bars 50, 57, and 58. Rouse is behind the beat in bar 50, and the delayed second note sounds as an
anticipation of beat 3 rather than 4 with respect to the rhythm section. In bar 57, he is behind the beat so far
that the delayed E♭ is simultaneous with a perceived beat 3 in the bass and drums. Of similar consequence in
bar 58, the A–F# notes are a 5:1 ratio, with the former heard on top of the beat and the latter perceived with
12
beat 3 in the drums and bass. Bar 90 is heard as written: Rouse emphasizes the F# with a rough timbre to
anticipate beat 3, con rmed in its metric reading with the note F in bar 91 and the descending interval from
beats 2–3 in bar 92. In sum, Rouse emphasizes beat 3 with the rst note of the cell in bar 45, and the
remaining occurrences emphasize a perceived beat 3 (rhythmically displaced or not) with the last note of
the cell through delayed microtiming or timbral in ection.

A full analysis of Rouse’s solo would include more examples of this cell. In fact, he uses it as a germ motive C19.P40
for phrases, sometimes transformed and augmented as in bars 60 and 91–92 shown in formula β. One
extension of the cell constitutes a formulaic family γ in Figure 19.13.

An extension of the core cell (γ.1) gives rise to the core of formulas γ.2 and its inverse γ.3. For each C19.P41
occurrence, the core is bracketed with its attendant number written above the phrase. In the examples that
include section markings of A, it is apparent that Rouse used this formula at similar times in the form,
notably at the beginning of the solo (bar 41) and to begin four of the remaining seven A sections. Limiting to
the speci c pitches of D, E♭, and F, this list of the formula reads as though the cores begin on di erent beats;
however, the core typically sounds as if it begins on a strong beat (1 or 3). Two examples are an exception.
Bars 96–98 and 72–73 are similar in notation: the continuation of the line resolves to the note B♭ on a
13
strong beat. Bars 72–73 are placed at the bottom to highlight this di erence, the nal note, B♭, con rming
that a beat has been inserted into the form. It is at that point that Rouse’s playing continues to be relatively
congruent with the metre implied by the bass and drums.

Rouse’s playing suggests that rhythmic displacements were part of his plan all along. He was a seasoned C19.P42
jazz performer, and many of his solos demonstrate that he could ‘play the changes’. In addition, he plays
similar gures in other recordings of ‘Evidence’ that do not contain metrical shifts. His style expanded on

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Monk’s musical conception, and one can consider the formulas to re ect the thematic material—a method
p. 405 for creating interest through rhythmic displacement. In the Blackhawk recording, however, the formulas
also contribute to a sense of metrical shift as shown next in a description of the group’s interaction.

Figure 19.13

Formula γ. Formulaic family of the melodic cell (γ.1 from formula β), and its extended use (γ.2 and γ.3). C19.F13

Interacting with Metre C19.S5

The analysis of individual parts above de nes the di erent signals that contribute to our perception of the C19.P43
p. 406 metre: Monk plays the melody on piano, Higgins provides comping cues on the snare and bass drums
while maintaining a displaced swing groove, and Rouse improvises with formulas that are sometimes
rhythmically displaced. John Ore contributes to the interaction by providing a walking bass line with the
harmonic rhythm indicated by 5–1 or chromatic note movement. The following analysis aims to highlight
the signi cant moments where the bass contributes to a sense of metrical shift. The harmonic rhythm
supports changes between the original and shifted metric paradigms, which o ers signals to the group as
part of the interaction.
The beginning of the saxophone solo involves a direct metrical dissonance. The bass provides a quarter- C19.P44
note pulse beginning on beat 2, and aligns a harmonic rhythm with the displaced swing pattern in the
drums. Monk’s whole-note chords are played against that metric reading. His emphasis of the downbeat is
supported by Rouse with formula γ and sparse ideas such as the melodic cell of formula β.

Six bars into the A section, the bass leaves the displaced swing pattern to align with the piano and C19.P45
emphasize the harmonic rhythm on beats 1 and 3. Soon after, the bars in Figure 19.14 demonstrate an

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
interplay that leads to a perceived metrical shift, and a local consonance between all parts.

Each instrument with its bracketed pair includes a metric reading according to the original and shifted C19.P46
metric paradigms (the top and bottom staves, respectively), and the musical materials discussed earlier. I do
14
not suggest that a sense of polymetre exists in these passages; rather, the arrows represent possible times
for hearing a metrical shift depending on which instrument one focuses their attention. I maintain that a
perceived shift of the metre is heard somewhere in these bars. For listeners, the indirect dissonance (the
point of switching between original and shifted paradigms) may arise somewhere in the beats between
arrows of the di erent parts. For the musicians, the process may have started earlier, or been attended to
after hearing the change. The following gures aim to explain why the shifts occur.

At the end of bar 48, the D–E♭ in the bass indicates the chord change across the bar line shown in the C19.P47
bottom bass stave. The strong chromatic movement across the remaining bar lines is annotated with
enclosures. Aligned with a standard swing groove shown in the bottom stave of the drums (with a snare-
drum comping pattern anticipating the downbeat of each bar), the saxophone presents a shift with the core
of formula γ.2 at the beginning of bar 49. The following melodic cell of formula β has a delayed E♭, sounding
as an anticipation of beat 3 and con rming the metrical shift. The sparse melody heard in the piano sways
the listener to hear the metre similarly; however, the silence in the saxophone brings attention to the
melody on beat 2 of bar 51.

A continuous reading is required between Figures 19.14 and 19.15. From bars 51–52, a metrical shift back to C19.P48
the original paradigm is occurs with the A–B♭ in the bass, and the ghost note of B♭ as an anacrusis to C in the
saxophone. The result is a momentary direct dissonance with the drums.

Transcription cannot capture the complexity of this passage. The C–E in the saxophone is a quote of the C19.P49
melody played a beat early. The E is normally heard on beat 3, and in this case, is shifted backwards to sound
as the downbeat of bar 52. Monk restates the melody an eighth-note rest early (the rhythm is on beats 2–3
p. 407 in the head). Those notes are heard to shift as well, followed by the bass with two B♭ notes (a tritone
substitution of the E7 harmony) perceived as beats 3 and 4. The metric reading is veri ed in the saxophone
from bars 52–53, with formula α resolving to G over the bar line. Monk’s melody on piano is heard as an
anticipation of beat 1 in bar 53, resulting in all instruments having local consonance with the standard swing
pattern in the drums. The metre is jolted back into place with the structural A♭m7–D♭7 chord changes in the
piano (bar 54). The root motion in the bass follows suit, and the 16th-note pattern in the saxophone is heard
beginning on the next downbeat: a direct metrical dissonance with the drums.
Figure 19.14

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Perceived metrical shi to a local consonance (the shi ed paradigm on the bottom staves of bracketed pairs), and back to a C19.F14
metrical dissonance in the original paradigm (top staves of bracketed pairs). Second A section of the AABA form, chorus 1 of the
saxophone solo.

Figure 19.16 begins with a continuation of the original metric paradigm, completing the saxophone’s C19.P50
rhythmically dense passage, and with the melody on piano. The bass o sets the metre in bar 56 with the
B♭–E♭ as a V–I movement, prepared by preceding C–F as a cycle of fths (see bar 55 in Figure 19.15). The B♭,
however, is also the V of E♭ and may be heard as part of the harmony. A metric reading of the bass is clearer
at the end of bar 56 with the downward chromatic movement resolving to the root of B♭m7.

p. 408 With the drums now heard as a standard swing groove, the polyrhythm and melody on the piano is C19.P51
perceived as a metrical shift to a local consonance, much like the head arrangement. On saxophone, the
melodic cells of formula β fall in line: prepared by the previous occurrence, the delayed notes anticipate beat
3 on the bottom t.s. stave (cf. Figure 19.14). The accented C♭ sounds as a downbeat in bar 60, followed by the
descending interval in an eighth-note rhythm to anticipate beat 3 again.
Figure 19.15

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Continuation from Figure 19.14, with a metrical shi to a local consonance in chorus 1 of the saxophone solo (the shi ed C19.F15
paradigm on the bottom staves of bracketed pairs), and back to a metrical dissonance in the original paradigm (top staves of
bracketed pairs). Second A section of the AABA form, chorus 1 of the saxophone solo.

The shifted paradigm remains through the B section with the chord changes in the bass, and the displaced C19.P52
melody of the head on piano. The next A section shown in Figure 19.17 continues in the shifted paradigm.
With Monk’s restatement of the metrically ambiguous melody, the bass is perceived to change chords from
bar 64–65, and con rmed with a descending G♭–C in bars 66–67 followed by a cycle of fths resolving on
the root of B♭7. Each pitch in the bass is repeated and followed by a leap to stress the strong beats. On
saxophone, the core of formula γ.3 on the downbeat of bar 65 further contributes to hearing the shifted
metric paradigm.

p. 409 Bar 68 carries two considerations that demonstrate Rouse’s rhythmic play. First, the notes C–E are a quote C19.P53
of the melody (bar 4) with its original rhythm from beats 2–3 on the top stave; the melody is supported by
the same notes in the piano voicing. However, one may hear this as a rhythmic displacement in the shifted
paradigm. Bar 67 is heard as a shifted metre, with the saxophone’s accented downbeat, and melody notes G–
C across the bar line to emphasize beat 1 of bar 68. After the melodic quote of C–E, formula α rea rms the
shifted paradigm, maintaining the shift according to the bass and drums.
Figure 19.16

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Continuation from Figure 19.15: chorus 1 of the saxophone solo in the AABA form, a metrical shi to a local consonance in the B C19.F16
section (the shi ed paradigm on the bottom staves of bracketed pairs) akin to the B section of the head (cf. Figure 19.3).

These aural metric shifts demonstrate how the group reacted to each other. More ambiguous is Figure 19.18, C19.P54
where a listener may hear one of the two metric paradigms and possibly switch between them more than
once. The original paradigm involves hearing the piano with the chord changes A♭m7–D♭7 from bar 6 of the
head, continued with the melody, and supported by the bass in bar 70. The shifted paradigm is intended to
show an alternative metric reading according to the saxophone, aligned with a standard swing pattern in
the drums, and supported by the bass after bar 70. Taken together, the metrically dissonant tension is
resolved when the saxophone and bass insert a beat into the form (i.e. the changing time signatures).

The adjustment by the bass comes early, followed by γ.3 in the saxophone when the rst three notes sound C19.P55
p. 410 like a 2–7–1 resolution in E♭. It is impossible to know exactly when Rouse adjusted his sense of the metre.
The core of γ.3 suggests that he was continuing with the original paradigm, and only after completing the
phrase did he nd himself in synch with the bass and drums; alternatively, he may have intended the
change bars earlier, or come to the realization that it ‘worked’ while playing it. At the same time, the bass
drum after bar 72 clearly marks the beginning of the next chorus and has been notated as inserting a 1/4 bar
15
to align with the saxophone and bass.
Figure 19.17

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Continued hearing in the shi ed paradigm (bottom staves of bracketed pairs) in the transition from the B to final A section of the C19.F17
AABA form, chorus 1 of the saxophone solo.

But what about Monk? The saxophone, bass, and drums align according to a standard swing groove at the C19.P56
end of the rst chorus, which leaves Monk on his own. As shown in Figure 19.19, he stayed true to his metric
conviction by keeping the comping patterns xed on beats 1 and 3, with rhythms on the beat, or
conventional anticipations of them. His momentary silence in the consecutive 1/4 and 4/4 bars brings an
adjustment, and is veri ed by his comping gure to articulate the beginning of the next A section.

The solo continues on this metric path for the rest of the piece. Of note, however, is the rhythmic interplay C19.P57
in the transition to the following B section. Rouse initiates a metrical shift with formula α, and Monk plays
p. 411 the B-section melody akin to the shift heard
p. 412

in the head. The metre is brought back to its resting place with a standard swing groove, the bass asserting a
harmonic rhythm, and the saxophone following with formula β. The back-and-forth shift of the metre
seems to comment on the turmoil from the rst chorus, and its quick resolution shows the group’s
command of time.
Figure 19.18

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Hearing in the original or shi ed paradigms (top or bottom staves of bracketed pairs), and the insertion of a beat by the bass and C19.F18
saxophone to align with a standard swing groove in the drums. In the AABA form, transition from chorus 1 to chorus 2 of the
saxophone solo.

Figure 19.19

Metric adjustment by Monk on piano; the piano inserts a beat to align with the swing groove of the other parts in the transition C19.F19
from the first to second A section, chorus 2 of the saxophone solo.

Concluding Remarks C19.S6

Perception of the metre presented in this chapter accounts for a listening experience by considering C19.P58
separate instrumental parts, and the combination of their musical features as modes of interaction. As a
result, those interactive qualities demonstrate that the metre shifts back and forth between metrical
dissonances and local consonances in the original and shifted paradigms.

In this performance of ‘Evidence’, entrainment involves following a groove that changes depending on how C19.P59
the musicians independently imply the beat. When Monk, Rouse, and Ore adhere to the original metric
paradigm, Higgins’s displaced swing pattern on drums creates a direct dissonance within the groove. The
equilibrium changes at a point of indirect dissonance: metrical shifts occur when the bass (indicating
harmonic rhythm) and saxophone (implying strong beats with formulas and microtiming) align with the
swing pattern in the drums.
These points of change are highly dependent on the rhythm on piano and, when Monk restates the melody, C19.P60
the metrical structure of the head arrangement. The sparse melody at the beginning of the A sections
a ords the opportunity for the metre to shift paradigms during the solo when the saxophone, bass, and
drums are aligned. The structural moment analogous to bar 6 of the head emerges as a signpost: the chord
change of A♭m7–D♭7 on beats 1 and 3 is a strong metrical signal that is followed in the bass and saxophone
to shift the metre back to its original paradigm. The B section—with rhythmically displaced chords from the
head arrangement in the piano—again a ords a shift in the metric paradigm: each melody note sounds like

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
an anticipation of the downbeat. The improvised lines in the saxophone and changes in harmonic rhythm in
the bass con rm the perceived shift.

Monk’s comping is not the determining factor for the shifts in and out of the original metric paradigm. The C19.P61
overturned swing beat and comping patterns in the drums, the saxophone’s formulas with microtiming and
solo lines according to chord changes, and the harmonic rhythm in the bass all play key roles in the process.
If the displaced swing beat in the drums at the end of the head arrangement was a mistake, the interactive
choices by everyone thereafter may be seen as attempted saves. In turn, the group members continually
interpret each others’ attempted saves as metrical signals during the improvised communicative process.

From this analysis, one can see how rhythmic displacements in the tune lend themselves to improvisation C19.P62
with respect to the metre. Whether Monk approved of the performance or not is subject to debate. Setting
p. 413 aside critical judgements about the performance and thinking about how it is experienced leads to an
extraordinary listening to a musical conversation with time.

Notes
1. See Love (2013) and his references within. C19.N1

2. See Koch (1983: 76–78), McLaughlin (1983: 86–87), and Wilson (1987: 42–46). C19.N2

3. Love (2013: 50–57) describes Mirkaʼs work as ʻechoingʼ London (2004) with its attention to entrainment while employing C19.N3
Hastyʼs theory of projection (1997). See Mirka (2009: 17–30) for her discussion of theory and method.

4. For the performance practice of playing ahead, on top, and behind the beat, see Berliner (1994: 146–147: 150–151: 245: C19.N4
351–352: 396: 413), Keil (1995: 8), and Monson (1996: 56).

5. Givan (2016) analyses musical features at the level of micro, macro, and motivic interaction, though not with a focus on C19.N5
musical time. The analysis of ʻEvidenceʼ in this chapter requires these levels of detail as they relate to rhythm and metre.

6. The concept of a separate metric paradigm notated on a stave below the original paradigm is borrowed from Haywood C19.N6
(1995).

7. A full transcription is available in Bruce (2013: 443, 445–457). C19.N7

8. See Berliner (1994: 617) and Monson (1996: 52–54). C19.N8

9. Monkʼs comping references the melody in many of his recordings. Solis provides a similar analysis on another recording of C19.N9
ʻEvidenceʼ (2001: 59).

10. Monk (1994), (2006: tracks 2 and 8), and (2007). C19.N10

11. See Spring (1990) for the isolation of a ʻcoreʼ in formulas. C19.N11

12. The last notes in bars 57 and 58 are within 50ms of the next beat, which is under the ≈100ms threshold of perceivable C19.N12
isochronous rhythmic attacks (London 2004: 27–29, 72). Perceiving the delayed ʻcoreʼ to begin on the backbeat may also
be partially explained as typical microtiming (Iyer 2002: 405–407).

13. These two examples are di erent from bar 89, when the resolution comes from the descending leap ending on beat 2. C19.N13
14. See London (2004: 49–50) for theory against the possibility of polymetre. C19.N14

15. The 1/4 bars are not included in the numbering, and the 1/4 and 2/4 bars in the bass are counted as one for the purpose of C19.N15
number alignment.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
References C19.S7

Berliner, P. (1994). (1994). Thinking in jazz: The infinite art of improvisation. University of Chicago Press. C19.P63
Google Scholar Google Preview WorldCat COPAC

Bruce, R. (2013). Change of the ʻguardʼ: Charlie Rouse, Steve Lacy and the music of Thelonious Monk. Doctoral dissertation, York C19.P64
University.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Collier, G. L., and Collier, J. L. (eds) (2002). Music Perception: An Interdisciplinary Journal, 19(3). C19.P65
Google Scholar Google Preview WorldCat COPAC

Do man, M. (2013). Groove: Temporality, awareness, and the feeling of entrainment in jazz performance. In M. Clayton, B. Dueck, C19.P66
and L. Leante (eds), Experience and meaning in music performance, 62–85. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Feurzeig, D. K. (1997). Making the Right Mistakes: James P. Johnson, Thelonious Monk, and the Trickster Aesthetic. D.M.A Thesis C19.P67
(Cornell University).
Google Scholar Google Preview WorldCat COPAC

p. 414 Givan, B. (2016). Rethinking interaction in jazz improvisation. Music Theory Online 22(3): 22.3.7. C19.P68
http://mtosmt.org/issues/mto.16.22.3/mto.16.22.3.givan.html.
Google Scholar WorldCat

Gushee, L. (1991). Lester Youngʼs ʻShoe shine boyʼ. In L. Porter (ed.), A Lester Young reader, 224–254. Smithsonian Institution C19.P69
Press.
Google Scholar Google Preview WorldCat COPAC

Hasty, C. (1997). Meter as rhythm. Oxford University Press. C19.P70


Google Scholar Google Preview WorldCat COPAC

Haywood, M. S. (1995). Rhythmic readings in Thelonious Monk. Annual Review of Jazz Studies 7: 25–45. C19.P71
Google Scholar WorldCat

Hodson, R. (2007). Interaction, improvisation, and interplay in jazz. Routledge. C19.P72


Google Scholar Google Preview WorldCat COPAC

Iyer, V. (2002). Embodied mind, situated cognition, and expressive microtiming in African-American music. Music Perception C19.P73
19(3): 387–414.
Google Scholar WorldCat

Keil, C. (1995). The theory of participatory discrepancies: A progress report. Ethnomusicology 39(1): 1–19. C19.P74
Google Scholar WorldCat

Koch, L. (1983). Thelonious Monk: Compositional techniques. Annual Review of Jazz Studies 2: 67–80. C19.P75
Google Scholar WorldCat

Krebs, H. (1999). Fantasy pieces: Metrical dissonance in the music of Robert Schumann. Oxford University Press. C19.P76
Google Scholar Google Preview WorldCat COPAC

London, J. (2004). Hearing in time: Psychological aspects of musical meter. Oxford University Press. C19.P77
Google Scholar Google Preview WorldCat COPAC

Lord, A. B. (1960). The singer of tales. Harvard University Press. C19.P78


Google Scholar Google Preview WorldCat COPAC

Love, S. (2013). Subliminal dissonance or ʻconsonanceʼ? Two views of jazz meter. Music Theory Spectrum 35(1): 48–61. C19.P79
Google Scholar WorldCat

McLaughlin, M. (1983). African music, rhythm and jazz. Proceedings of NAJE Research 3: 74–91. C19.P80
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470722 by National Science & Technology Library user on 26 May 2023
Mirka, D. (2009). Metric manipulations in Haydn and Mozart: Chamber music for strings, 1787–1791. Oxford University Press. C19.P81
Google Scholar Google Preview WorldCat COPAC

Monk, T. (1987). Thelonious Monk Quartet Plus Two at the Blackhawk. Riverside OJCCD-305-2, CD. Recorded 29 Apr. 1960. C19.P82
Google Scholar Google Preview WorldCat COPAC

Monk, T. (1994). Thelonious Monk Big Band and Quartet in concert. Columbia C2K 57636, CK 57849, CK 57850, 2 CD. Recorded 30 C19.P83
Dec. 1964.
Google Scholar Google Preview WorldCat COPAC

Monk, T. (2006). Thelonious Monk: In Philadelphia 1960 with Steve Lacy. Rare Live Recordings RLR88623, CD. Recorded 16 Feb. C19.P84
1948, June 1950, 6 Oct. 1955, Aug. 1957, and 3 Mar. 1960.
Google Scholar Google Preview WorldCat COPAC

Monk, T. (2007). Live at the 1964 Monterey Jazz Festival. Monterey Jazz Festival Records MJFR-30312, CD. Recorded 20 Sept. 1964. C19.P85
Google Scholar Google Preview WorldCat COPAC

Monson, I. (1996). Saying something: Jazz improvisation and interaction. University of Chicago Press. C19.P86
Google Scholar Google Preview WorldCat COPAC

Solis, G. (2001). Monkʼs music and the making of a legacy. Doctoral dissertation, Washington University. C19.P87
Google Scholar Google Preview WorldCat COPAC

Spring, H. 1990. The use of formulas in the improvisations of Charlie Christian. Jazzforschung/Jazz Research 22: 11–51. C19.P88
Google Scholar WorldCat

Wilson, P.N. (1987). (1987). Versuch über das ʻMonkischʼ: Zur musikalischen Ästhetik Thelonious Sphere Monks und ihrem C19.P89
posthumen Weiterwirken. Jazzforschung 19: 41–59.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
CHAPTER

20 The Politics ofC20Musical Time in the Everyday Life of


Ballet Dancers 
Jonathan Still

https://doi.org/10.1093/oxfordhb/9780190947279.013.20 Pages 415–442


Published: 08 December 2021

Abstract
This chapter discusses disagreements and misunderstandings about musical time in the context of
ballet classes and rehearsals, and the degree to which musicians’ metric-counting is regarded by both
musicians and dancers as more correct than ‘dancers’ counts’. Metrical anomalies in music by
Tchaikovsky, Bizet, and Verdi used in children’s ballet classes are examined in the light of research by
William Rothstein on national metrical types and Franco-Italian hypermetre, and found to be less
anomalous than they might seem at rst. The problems of representation and human movement in
these examples are discussed with reference to debates about dance in non-representational theory
(NRT), and conceptual and disciplinary boundaries in music and dance scholarship.

Keywords: musical listening, music in everyday life, ballet training, music education, metre, rhythm,
choreomusicology
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

THE story [in the song and dance lms] was the price you paid for the rhythm. ‘Pardon me, boy, is C20.P1
that the Chattanooga choo choo?’ Each syllable found its corresponding movement in the legs, the
stomach, the backside, the feet. In ballet hour, by contrast, we danced to classical recordings
—‘white music’ as Tracey bluntly called it—which Miss Isabel recorded from the radio on to a
series of cassettes. But I could barely recognize it as music, it had no time signature that I could
hear, and although Miss Isabel tried to help us, shouting out the beats of each bar, I could never
relate these numbers in any way to the sea of melody that came over me from the violins or the
crashing thump of a brass section. (Smith 2017: 24)

With Ondine, from the outset, we could hear the colour texture and emotional force of the music C20.P2
which not only conveyed the atmosphere of the drama; it was to produce an unusual homogeniety
[sic] of music and dance. But, at rst, the dance rhythms seemed inextricably embedded in the rich
and complicated scoring. Here the help of the rehearsal pianist was of the utmost importance, as
he alone had the knowledge to translate the apparently rhythmless ow of sound into countable
phrases.

(Hynd 1961: 28)

Introduction C20.S1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
p. 416

The passages cited above describe almost the same musical problem experienced in vastly di erent settings. C20.P3
In the rst, from Zadie Smith’s novel Swing Time (2017), the unnamed narrator is recalling the Saturday
morning tap and ballet classes that she took as a 7-year-old child in London. In the second, the former
Royal Ballet principal Ronald Hynd, aged 30 at the time, is writing about the di culties encountered when
working with Hans Werner Henze’s music for Ondine (1958), commissioned and choreographed by Sir
Frederick Ashton.

These rare examples of everyday musical confusions in the ballet world made public illustrate the main C20.P4
topics of this chapter: rst, the practice, in children’s ballet classes as well as professional rehearsals, of
teachers or choreographers mediating between dancers and music with systems of ‘dancers’ counts’;
second, the sometimes problematic and enigmatic relationship between these counts, time signature, beats,
and metre; and third, the tension between the idea of a direct, unmediated experience of music, and one in
which forms of representation are involved. It is easy to see why Ashton might have needed a pianist to help
him and his dancers make sense of Henze’s music, but I will argue here that there is every reason to pay
more attention to problems experienced with apparently much simpler music in children’s ballet classes.

If there is a reason why problems like this attract little attention in either music or dance education C20.P5
research, it might be partly because a child’s inability to recognize what she heard as music, or to discern its
time signature, is likely to be viewed as a lack of musical aptitude or knowledge, something to be corrected
rather than explored. Time signature, and its associated conventions in music pedagogy of metrically
weighted counting schemes (i.e. counting 3/4 as 1 2 3 1 2 3 etc.), are often assumed to constitute a technical,
logical, perhaps even mathematical standard, against which the ‘dancers’ counts’ (or ‘dance counts’) that
are commonly used in ballet rehearsals and classes are what could be called ‘folk’ counting: a practical but
ultimately not very musical method that gets the job of teaching or learning choreography done.

One way to describe dancers’ counts is as a temporal framework for connecting music and dance—a C20.P6
numbered timeline of key moments in a movement sequence that also aligns with the most salient pulses in
the music as heard (or that will align with them in the case of choreography that precedes the selection or
composition of music), but not necessarily at the metrical level suggested by the time signature (Jordan
2000: 81; Still 2015). Counts may also sometimes cut across or ignore what musicians would regard as the
metrical or phrasal structure of the music, so that in a rehearsal musicians and dancers might ask questions
1
like ‘How are you counting that?’ or ‘Where’s your “one”?’ Ballet is perhaps more prone to problems of
coordination between movement and music than other dance forms, due to its frequent pairing with
Western art music, in which metre is an organizing principle, but not necessarily a prominent surface
feature of the sound.

p. 417 But even in the apparently more accessible nineteenth-century repertoire, metrical structure may be C20.P7
ambiguous, and sometimes deliberately so. The coda of the pas de trois (Act I No. 4: VI) in Tchaikovsky’s
Swan Lake Op. 20 (1875–76) (see Figure 20.1) could be counted as either two or four in a bar, and bar 2 can be
treated either as the rst bar of a melody with an empty downbeat, or as an anacrusis to a ‘one’ at the
beginning of bar 3. When I asked about a dozen professional dancers or coaches familiar with the
2
choreography most commonly performed to this section of music, three possible counting schemes
emerged, none of which concurs, I think, with how musicians would interpret this passage metrically.
Regardless of di erences in counting, all three versions place the dancer’s ‘1’ in the middle of the second
bar, not at the beginning of either bars 1 or 2. As strange as version A may seem, there is nothing to stop a
‘notional 1’ being on count 2. In the choreography dancers referred to, this beat marks the take-o for a
3
jump, so it would be a case of saying that you ‘jump on 2’. For musicians used to the idea of counts always
carrying metrical weight, this is confusing.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Figure 20.1

Tchaikovsky, Swan Lake Op. 20 (1875–6), Act I, Scene 4: VI, Coda from the Pas de trois. C20.F1

With experience, musicians working in ballet soon begin to adopt the terminology of counts rather than C20.P8
bars, marking up their scores showing the structure of the dancers’ counts, so that they can see quickly
where to start playing if a coach suddenly asks: ‘Give us four into the sixth eight’ or ‘Play one eight before
her arabesque’. Dancers—who, unlike singers or instrumentalists, are never working from a score—often
make light-hearted apologies when talking to rehearsal pianists about music such as ‘You’ll probably hate
the way I’m counting this, but for me it’s two sixes and a four’, or ‘I’m probably counting this all wrong but
that’s how I hear it’. Choreographers and dancers have di erent opinions about counting: how and when it
should be done, for how long, or even whether to do it all, and to what extent it helps or hinders
performance and creativity (Jordan 2000: 93–95, 187–188, 279; Kossen-Veenhuis 2017: 93–100).

So far, so unpolitical. Negotiating the di erences between counts and notation is simply part of everyday C20.P9
working practice in the ballet world, even if, privately, some dancers and musicians nd it annoying to have
to count in ways that they nd counterintuitive. However, for ballet dancers who are making the transition
p. 418 to becoming teachers, the situation is rather di erent. A grounding in music theory, albeit sometimes
basic and brief, is still common on teacher training courses, and even though most teachers outside of
vocational schools and professional dance companies work with recorded music, it is common to hear
variants of the following view on the teachers’ need for music theory: ‘Being musically knowledgeable is a
necessity for a ballet teacher. A fundamental knowledge of music aids the teaching process in many ways;
[…] 1) in communicating with an accompanist in ‘musical’ as opposed to ‘dance’ terms, which usually seem
rather vague to a musician’ (Warren 1989: 73).

In practice, what a ‘fundamental knowledge of music’ means is learning how to count music ‘properly’, that C20.P10
is, according to the conventions of Western metric notation. Those who teach music theory on ballet teacher
training courses are usually musicians, and as a result, they are prone what Jeanne Bamberger calls a ‘wipe-
out phenomenon’—an erasure of the memory of what it was like to perceive music before they could read
music notation, because they have internalized the rules of metric notation to the extent that hearing and
4
thinking in those terms becomes second nature (Bamberger 2013: 40–41, 252–254). As a result, musicians
are prone to confuse what they know about music with what they perceive—that is, so that what they can
say about music becomes simply ‘knowledge of music’. The problematic educational implications of this
phenomenon are that metric-formal descriptions can be regarded by music teachers as what appear to them
to be a naturally occurring ‘right’ answer, in comparison to what are perceived to be the wrong or ‘less
adequate’ gural hearings of the so-called ‘naïve listener’ (Bamberger 2013: 40). The Swing Time narrator’s
comment that the music in her ballet class had ‘no time signature that [she] could hear’ exempli es the
issue: you cannot ‘hear’ time signature, time signature is not part of the musical sound, but it is not at all
unusual to speak in those terms.

Bamberger’s distinction between ‘units of description’ (the language, symbols, and analytical categories C20.P11
that musicians use to talk to each other about making music) and ‘units of perception’ (how untrained

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
listeners or self-taught musicians make sense of the music they hear) corresponds, roughly speaking, to
5
Philip Tagg’s terms ‘poïetic’ and ‘aesthesic’ (Tagg 2012: 116), or Nicholas Cook’s di erentiation between
‘musicological’ and ‘musical’ listening, where the purpose of the former is ‘the establishment of facts or the
formulation of theories’, and the latter ‘for purposes of direct aesthetic grati cation’ (Cook 1992: 152) that
has a ‘non-dualistic quality’ (p. 153). Musicological listening, typi ed by what Virgil Thomson called ‘the
Appreciation-racket’ (Thomson 1939, cited in Cook 1992: 69, 163–164), is characterized by a focus on the
productional categories of music understood by musicians and theorists, and requires listeners to re ect
consciously on formal elements and structures—an activity that threatens to get in the way of being able to
experience and enjoy music spontaneously or directly, since the process of listening is interrupted or even
replaced by re ection. However, even for those with musical training, there are ‘striking discrepancies
between the structure of [listeners’] experiences and the structure of the music itself, as one would analyse
it on the basis of the score’ (Cook 1992: 58). If this is true of a sedentary listener, it is also pertinent to the
p. 419 experience of a moving dancer, for whom listening attentively for compositional processes in music
would be an irrelevant distraction from the demands of dancing, especially given that choreographers
devise movements in sympathy with a particular hearing and interpretation of music that might di er from
one grounded in conventional music analysis, or use music to underscore expressive action in a way that is
only loosely bound to compositional structure.

In this chapter, I follow Bamberger’s lead in trying to make sense of what might seem to a musician to be C20.P12
anomalies in dancers’ counting schemes, not just to understand them, or to dismiss the value of notation or
attentive listening, but to draw attention to anomalies and variations in the counting conventions and
practices of music notation. These may be familiar territory for academic music theorists, but they are
rarely if ever brought to bear on everyday musical practice in the ballet world, where they have a surprising
relevance and applicability. As an occasional teacher of music theory to dance teachers and an accompanist
for schools and ballet companies myself, I have a longstanding interest in such matters, but it was in my
role as an editor for music publications to accompany ballet examinations that these matters became most
intriguing and urgent, since they involved making di cult editorial decisions about anomalies of notation,
that would make an orthodoxy of a phenomenon that I was confused by myself. What was di cult about
these particular examples was the apparent illogicality of the original music notation, rather than the
dancer’s counting of the music. In the second part of the chapter, with reference to the work of William
Rothstein (2008; 2011), I show how an understanding of Franco-Italian notational conventions in the mid-
nineteenth century helps to illuminate and explain these anomalies. This expands the scope of Rothstein’s
argument to include not just the historical ballet repertoire, where there are several examples of the
6
principles he describes in addition to the ones I discuss here, but everyday embodied practice in ballet
classes and rehearsals, where this repertoire and its metrical peculiarities are encountered by many
musicians, teachers, and pupils. In the remainder of this rst part of the chapter, I outline some of the
intellectual background to these ideas.
Research into Music in Ballet Classes: An Interdisciplinary Gap C20.S2

The absence of research into ballet teachers’ everyday engagements with music in the ballet world is rather C20.P13
remarkable. Dance historians, musicologists, and choreomusicologists are mostly concerned with works by
7
high pro le composers, choreographers, or composers, as performed by adult professionals, whereas
studies of music in ballet training, in children’s ballet classes, and in other non-professional contexts are

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
almost nonexistent. This is not to imply that choreomusicologists ought to take up these challenges, but that
8
those one might expect to have done so, such as dance educationalists and ballet pedagogues, have not. In a

p. 420
children’s ballet class, issues of music notation and works of music or dance do not oat in a symbolic
domain outside everyday life; they are part of it. As such, they deserve to be taken more seriously as topics in
dance education research. In the examination syllabi produced by dance teaching and awarding bodies such
as the Imperial Society of Teachers of Dance (ISTD) and Royal Academy of Dance (RAD), getting the correct
9
coordination of steps and music in an exercise or dance is part of what is being assessed, and in this sense,
those syllabi constitute collections of miniature works. The RAD, for example, has published most of its
examination syllabus materials in the form of audio and video recordings, printed piano music, ‘word
notes’ (written descriptions of the exercises and dances, with counts that indicate their temporal alignment
10
with the music), and Benesh Movement Notation (BMN). The exercises and dances may be small in scale,
and intended for study and examinations rather than performance on stage, but this does not necessarily
disqualify them from consideration as works (Goehr 1992: 244). In view of their multiple forms of
representation and speci cations, they arguably qualify as ‘thick works’ to use Stephen Davies’s (2003: 39)
phrase—highly prescriptive in their instructions to performers, in contrast to ‘thin’ ones which leave some
parameters open to choice and improvisation. There is as yet a gap in both music and dance education
11
research where the material realities of music in children’s ballet classes might sit. By ‘material reality’ I
mean answers to questions like ‘What do teachers and their students actually do with music in a class? What
music do they use? How and why do they use it? How and where is it produced, and by whom? Why this
music and not that? How or when do they count or not count? In what ways do they mediate, if at all,
between the music and their students, using words, counts, and gestures? Is one kind of counting regarded
as better than another? Is it better not to count at all?’

Although it is tempting to see this gap being closed by interdisciplinary research in music and dance, a lot C20.P14
12
hangs on what Born (2010: 210–213) refers to as the ‘modes’ and ‘logics’ of interdisciplinarity: how and
why it is done. In practice, almost by de nition, dance education as a discipline excludes music as either
13
irrelevant to its primary research aims, or as something to be left to the music department. If any of Born’s
modes of interdisciplinarity can be understood to apply here, it would I think be the ‘subordination-service’
kind, where music is co-opted into dance teacher training as a service discipline.

However, in my view, a more tting way to understand the relationship between the current C20.P15
(mis)application of music theory to ballet practice is Paul Dowling’s critique of school mathematics
textbooks designed for lower-ability pupils (Dowling 1995; 1998), where domestic settings such as cooking,
DIY, or shopping are recruited as means of making mathematics relevant to everyday life. In the process, the
distinctiveness of both mathematics and everyday life su ers: elements of shopping are disembedded from
their material setting (which is only partially described) in order to make a mathematical point.
Mathematics, meanwhile, appears less as a practice in its own right than as something which refers to (and
is thus necessary as a tool for participation in) everyday life, which Dowling calls ‘the myth of reference’
and ‘the myth of participation’. Even more pertinent is his notion of ‘the myth of certainty’, whereby
p. 421 ‘mathematical practice is represented by elementary arithmetic which constitutes the essential truth
residing at the heart of the universe’ (1998: 295). In the cases I describe later on, the quasi-arithmetic
certainty of elementary metrical theory will not su ce to explain the anomalies that can be experienced in
even quite simple music. However, these examples are not usually brought into discussion, precisely
because they break the rules that they should exemplify. Unlike its academic counterpart, music theory in
ballet teaching is not subject to debate, speculation, and investigation, but used, much pared-down, as a
tool.

The kind of bewilderment that the narrator of Swing Time experiences in her ballet class is thus likely to be C20.P16
diagnosed by ballet teachers or musicians as a symptom of being ‘unmusical’, or not yet having understood
how to count. Both judgements may turn out to be warranted, but neither of them allows for the possibility

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
that there might be something strange about the music or the teacher’s counting, or that not counting at all
might have made it easier for the pupil, or that it is in fact the pupil’s prior knowledge and understanding of
music and other dance styles that has confused her. In an ethnomethodological sense, one can regard the
confusion about music like this as a ‘breakdown’ (DeNora 2000: 89–90) that reveals something about the
everyday lived order, and consider musical representation as a topic, rather than a resource, for analysis
(Pollner and Emerson 2001: 121).

Another reason why everyday musical problems like this do not receive much attention in ballet pedagogy is C20.P17
that music is popularly perceived as being at its most powerful and magni cent when its e ects are felt to
be immediate and inexplicable, and so it is, so to speak, a badge of honour when no theory, thought—or,
heaven forbid, counting—appears to intervene between listeners and their experience of music. Similar
romantic notions of dance pervade literature in other disciplines—as a symbol of wordless abandon,
freedom, play, and ecstasy, being caught up in the moment. Ballet, with all its rules, discipline, foursquare
music, exercises and canon of works, not to mention its perceived elitism, has little of this allure. In
contemporary dance, autonomy from musical structures has been valued as one of the emancipating
innovations of the modern dance pioneers of the early twentieth century. Ballet, on the other hand, is often
perceived, as dance critic Rosalyn Sulcas (2013) has put it, to be ‘stuck […] in its expressive, or dependent
relationship to music’.

How such contrasting values might be expressed or reproduced in dance training is suggested by Sylvia C20.P18
Faure (2000) in her socio-anthropological study of dance teaching in France. Faure observed an emphasis
on counting as a distinguishing feature of ballet classes compared to contemporary dance. In ballet classes,
14
the musical structure of the dance movements was ‘objecti ed’ by the teacher through a framework of
counts (which she regards as themselves objecti cations of the beats of the music) that reduce the
15
possibility for dancers’ own interpretations (Faure 2000: 136). In contemporary dance classes, by contrast,
students were encouraged to respond in personal, divergent ways to their hearing of the music, with the
teacher acting in a coordinating role. She compares these di erences between musical practice in ballet and
16
contemporary dance classes to Bamberger’s distinction between ‘metric’ and ‘ gural’ hearing
p. 422 respectively (2000: 133–134). Although Faure is reporting only on what she observed in particular classes
in France, the association of ballet with the metric and with top-down choreography, and contemporary
dance with the gural and with bottom-up creativity, re ects a wider prejudice about the inferior status of
learning a dance that someone else has choreographed, compared to creative improvisation and dance-
making: the use of the term ‘creative’ next to improvisation already presupposes that the alternative is
uncreative (see Taylor 2003: 166–172 for a particularly vigorous discussion of this point).

The prejudice extends to the practice of counting itself. In Rhythmanalysis (2004), for example, a text that C20.P19
has been taken up enthusiastically by some in contemporary dance scholarship, Henri Lefebvre says: ‘[W]e
tend to attribute to rhythms a mechanical overtone, brushing aside the organic aspect of rhythmed
movements. Musicians, who deal directly with rhythms, because they produce them, often reduce them to
the counting of beats [des mesures]: “One-two-three-one-two-three” ’ (Lefebvre 2004: 6; emphasis
original). In this binary opposition, ‘rhythm’ becomes a placeholder for all that is ambiguous, everyday,
embodied, organic, ine able, and beyond representation; the constitutive outside, so to speak, of music
theory, which is nothing but counting. Rather surprisingly, the strongest defence of the representational as
a part of the everyday in dance teaching comes not from within dance pedagogy, but in critiques of the
treatment of dance in non-representational theory (NRT), a strand of thinking in cultural geography
17
associated in particular with the work of Nigel Thrift, which is why I devote some space to those
arguments in what follows.

Non-representational Theory, Dance, and Everyday Life C20.S3

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
The attraction of dance for non-representational theorists like J. D. Dewsbury (2011), Nigel Thrift (1997), C20.P20
and Derek McCormack (2002; 2008; 2013) is that for them it o ers an a ective, expressive, processual,
playful, and rhythmic engagement with the world that is to some extent irreducible to, unconstrained by,
and incapable of representation. It is a potent example of the mobility and performativity of practice, an
ever-changing, moment-to-moment embodied engagement with a world that is also continually in ux,
and of a space that is produced through moving. Through any practice, but especially dance, body-subjects
may evade, resist, or exceed the terms of their own representation and social situatedness, and in another
sense, such practices also tend to lie outside, elude, or defy representation. If this sounds vague and
optimistic, it is because the theory itself is like that—‘wilfully restless in character […] and thus tricky to pin
down’, as Hayden Lorimer says (2005: 94) in the rst of three overviews of NRT which nonetheless attempt
to do just that (Lorimer 2005; 2007; 2008).

Some critics of NRT have taken issue with the tendency to conceive of dance as a virtual, ideal world that can C20.P21
be distanced from problems of representation. George Revill’s (2004) experiences of learning French folk
p. 423 dance as an adult beginner in Nottingham, for example, led him to challenge the notions of ‘dance as free
unre ected expression and music as unmediated communication’. In his view, the representational should
not be considered ‘as a distinctive sphere of experience’, but one which is ‘intricately connected into and
supported by semiotic systems and spatial practices operating simultaneously at a multiplicity of levels and
scales’ (Revill 2004: 208). Likewise, Catherine Nash argues: ‘Not only is dance always mediated by words as
it is taught, scripted, performed and watched but dance is also often highly formalized and stylized; even
untrained dance is culturally learnt and culturally located’ (2000: 658). Her criticisms draw on arguments
from dance studies, in which there are already strong objections to a ‘classically Romantic tradition of
desiring to return to an unmediated, authentic relationship to the world, to be like “primitive” others who
are unburdened by thought—dancing women, the “exotic people”, rural peasants, to reject the modern in
favour of the “primitive” ’ (2000: 657). These arguments resonate with Lucy Green’s critique (2008) of the
position taken by John Shepherd and colleagues (1977),—that notation is an obstacle to the kind of direct or
immediate experience of music enjoyed by members of pre-literate societies. Green argues that this
characterization fails to take into account that ‘all experience of music, pre-literate or literate, requires
familiarity with style and is therefore never “immediate” but mediated through conventional meaning and
learnt’ (Green 2008: 233). In her view, this simply exchanges one orthodoxy for another, valorizing the
‘immediate, total, sensual pre-notational musical experiences of pre-literate societies’, while attacking
‘the rei ed, dualistic, rationalist, notational musical hegemony of industrial humanity’ (p. 233). In such an
all-or-nothing dichotomy, the task of engaging critically with problems of music, dance, and notation
disappears.

Frédéric Pouillaude (2017) comments on a related tendency in philosophy to idealize dance as a form of C20.P22
ecstatic play, rather than write about particular works of dance. In the work of Nietzsche, Paul Valéry, Erwin
Straus, and Alan Badiou, he writes:

Dance (the rst letter becomes capitalised) is never considered positively as a real, empirical art C20.P23
form, with authors, works, venues, and speci c dates, but rather as a private bodily experience
that is simultaneously infra- and supra-artistic […] Not only do these thinkers never mention
speci c choreographic works in their texts, they also describe dance in a way that suggests that
Dance has nothing to do with producing or showing dance works: dancing they suggest, consists
mainly in expending time and energy, in the present, in the solipsism and hedonism of
kinaesthesia; dancing becomes a kind of orgasmic ecstasy without product or the possibility of
transmission.

(Pouillaude 2017: x–xi)

Dance works are ‘absent’ in a second sense, though, since what are commonly referred to as works in dance C20.P24

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
are characterized by what Pouillaude calls an ‘ontological weakness’. For example, although Swan Lake is
one of the most frequently cited examples of a dance work, there is no single, stable choreographic work to
which Swan Lake refers. In case it needs emphasizing, this is true of the underlying musical work as well:
p. 424 cuts, changes, and interpolations in ballet scores are not just a common occurrence—there is probably
not a production of Swan Lake in which they do not occur, or one which uses Tchaikovsky’s complete score
in the original order as published in 1877. In addition, Pouillaude argues, unlike painting or writing, dancing
is an activity that leaves no durable trace. As a result, in terms of Hannah Arendt’s distinction between
‘work’ and ‘labour’ in The Human Condition (1998), dance can be seen not as the diametric opposite of
labour, but as a form of unproductive labour itself (Pouillaude 2017: 60–61), hence his choice of the term
désoeuvrement, or ‘unworking’. Paradoxically, though, those very philosophers whom Pouillaude criticizes
for their lyrical abstractions about dance have nonetheless, in his view, partially grasped something true
and important about how choreographic practices ‘unwork’ the work, and about what it feels like to dance,
but without knowing from experience (unlike Pouillaude, who was a dancer himself) why they are right.

18
In a similar way, there is a kind of doubleness about NRT, in that on the one hand, its very name suggests C20.P25
that representation lies outside its area of interest, yet on the other, NRT also provides a way of thinking
about those bewildering aspects of everyday musical experience in ballet classes—including representation
19
—that are described by the narrator of Swing Time at the beginning of this chapter. The sticking point is
what counts as the everyday, as I discuss below.

The Everyday C20.S4

The idea of a domain of pre-re ective, immediate, spontaneous experience is sometimes denoted by the C20.P26
term ‘everyday’. Sociologist Norbert Elias (1998[1978]) was critical of the tendency amongst his
contemporaries to use ‘everyday’ to mean a multitude of things which were rarely de ned explicitly.
Instead, they gave the impression that they were either for or against some unspoken Other that was ‘not-
everyday’. As a result, to derive a de nition of ‘everyday’ from some of the sociological literature, one had
to work out from the context what this implied Other was. Two uses in particular troubled Elias, because in
20
his view they indicated that the existence of the everyday was more an article of faith than a useful
sociological concept. In one, ‘everyday’ denoted a sphere of naive, false experience, compared to what was
true, correct, and genuine. In the other, it implied the opposite: a ‘sphere of natural, spontaneous,
unre ecting, genuine experience and thinking’, as opposed to one of ‘re ective, arti cial, unspontaneous,
especially scienti c experience and thinking’ (p. 171). This suggested a romantic longing on the part of
scholars for an area of life blissfully free from the need to think, where one could experience the world
spontaneously—or, as Nash puts it, a ‘division of labour separating academics who think (especially about
not thinking or the noncognitive) and those “ordinary people” out there who just act’ (2000: 662).

In recent music scholarship, a focus on everyday life usually involves a turning away from a perceived C20.P27
previous overemphasis on musical works and the meanings or technical analyses assigned to them by
p. 425 theorists, and a concern instead with the meanings that music has for listeners (in the broadest sense of
21
that term and the uses they make of it). It is an enormous and heterogeneous eld that can include: a focus
on music as a ‘technology of the self’ for the aesthetic ordering of private space, for emotional self-
regulation, or in the search for and construction of self-identity, and conversely, how music is used in
public spaces such as shops, aerobics studios, or neonatal wards for collective social ordering (DeNora
2000); background music designed to be unobtrusive (Lanza 1995; Stockfelt 1997); music that tends to go
unnoticed, either by listeners generally or by scholars (Kassabian 2013a; 2013b; Quinones et al. 2013; Tagg
2011), or the psychology of di erent kinds of listening and engagement with music (Becker 2004; Biddle
2011; Clarke 2005; Clarke et al. 2010; Gritten 2011; Herbert 2011). It includes the eld of mobile music studies
pioneered by Michael Bull (2000; 2007) and celebrated in a two-volume Oxford Handbook on the topic

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
(Gopinath and Stanyek 2014; 2017). In turn, this research forms part of a wider eld of auditory culture and
sound studies, in which some representative works might include Sterne (2003; 2012), Bull and Back (2015),
and Schwartz (2011).

In my analysis of the musical examples in the second part of this chapter, what I mean by ‘everyday life’ is C20.P28
not what happens outside the studio, in dancers’ spare time, or everything except musical works or
notation, but literally, anything that happens in a given moment or episode during a class or rehearsal in
relation to the work of ballet teaching and music. In this chapter, that even includes the decision-making
involved in publishing a musical score. In using the term ‘everyday’ in this sense, I follow DeNora’s (2014)
22
example: like Elias, she rejects the idea that the everyday is an enclave removed from other social worlds,
or that there are places or activities that are more everyday than others, and questions whether there can be
a sociology ‘of’ such things in their own right. In her view, the everyday, in the sense of the routine, the dull,
or the repetitious, su uses life everywhere and all the time, including, for example, the seemingly exalted
pursuits of science and religious devotion. By the same token, the extraordinary or the numinous can
suddenly in ltrate or interrupt the most mundane tasks. Thus, the everyday does not signify a particular
place or kind of activity, but ‘the site where experience is made manifest, where it takes shape, where sense
is made’ (2014: xx). This is the spirit in which I discuss the examples below: it is in such encounters between
the notation, performance practice, and dance that sense is made of the ‘reality’ of metre and time
signature.

Example 1: Childrenʼs Galop from Tchaikovskyʼs The Nutcracker C20.S5

The Nutcracker Op. 71a, composed by Tchaikovsky in 1892, and choreographed in countless versions around C20.P29
the world ever since, is probably one of the most frequently performed ballets of the last 100 years. In part or
p. 426 whole, it is a staple of end-of-term shows at ballet schools at Christmas time, and many ballet
companies use children from local schools to play the parts of young guests at the party scene in the rst
act. Part of this scene, halfway through No. 3 (‘Children’s Galop, entrance of the parents’), is a jig-like tune
in 6/8 at a walking tempo, a dance rhythm that is also common in children’s ballet classes since it is suitable
music for galloping and skipping.

Figure 20.2 depicts the tune as Tchaikovsky notates it. I have sketched the main features of the C20.P30
accompanying gure in the rst two bars, which continues underneath the violin melody. Dancers,
particularly young ones, will need to nd a way of knowing when to start. They are usually told to ‘count
three in’, and then, on what musicians call the ‘upbeat’ or ‘pickup’, or ‘anacrusis’, they start to dance,
counting the anacrusis as ‘1’. If a pianist is present for the rehearsal, the teacher or choreographer might
ask: ‘How many counts in do I have?’ Most musicians, unless they are used to working in dance, will say
‘four’. The teacher might then ask, ‘So the music is in four, then?’ and the musician might respond, ‘No it’s
a 6/8’, then explain that this is a six, but is counted in two. An inexperienced musician will claim that the
dancers do not know how to count ‘properly’; a seasoned one will say ‘It’s three in’, knowing that what
matters is what helps the children, rather than what is in the score.
Figure 20.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Tchaikovsky, The Nutcracker Op. 71a (1892), Act 1, Scene 1 No. 3: ʻChildrenʼs Galopʼ and ʻEntry of the Parentsʼ. C20.F2

In conventional music theory 6/8 is de ned as a form of compound duple time, consisting of two beats, a C20.P31
C20.P32
strong and a weak, subdivided by three. Dancers who count the main beats of the introduction as ‘1, 2, 3’ and
then start counting from ‘1’ again on the ‘weak’ half of the second bar will therefore be thought of—or think
of themselves—as ‘unmusical’, or just plain wrong. To be consistent with the musician’s view of metre,
dancers would have to count ‘8 and 1’ if their movement starts on the anacrusis. For the dancer, however,
this is impractical. If you have to do eight identical skips to the left, and then eight to the right, it is di cult
and rather pointless to try to force yourself to count your rst step as ‘8’ and your second as ‘1’.

So why is Tchaikovsky’s introduction three counts, rather than four? If, as far as a listener is concerned, the C20.P33
rst beat of the tune starts on the second half of the second bar, why not just shift it forward half a bar? Is
p. 427 anything gained by playing what appears visually as an anacrusis as if it occurred on a weak beat? Part of
the answer lies in the source of the tune, which Tchaikovsky borrowed from a French song, ‘Monsieur
Dumollet’, the melody of which dates from 1809 and can be found as no. 866 in Capelle (1811: 370). Seeing
the song’s lyrics make it easier to understand why the tune is barred as it is: the end-accented patterns of
French verse suggest if not demand it (see Figure 20.3).

Figure 20.3

ʻMonsieur Dumolletʼ, the French song on which Tchaikovskyʼs melody is based. C20.F3

Figure 20.4

ʻThe steamboat quickstepʼ from Howe (1843). In this version, the melody is shi ed to the first beat of the bar. C20.F4

The same melody, with minor di erences, appears about 30 years later as ‘The steamboat quickstep’ in an C20.P34
American tune-book for the accordion (Howe Jr, 1843) (see Figure 20.4). In this version, however, the
beginning of the tune has now been shifted to the rst beat of the bar. A comparison between these two
versions of the same tune illustrates the distinction Rothstein (2008) makes between Franco-Italian and
German metre. The di erence is already present, in fact, in contrasting theories of metre among
eighteenth-century theorists; but on the basis of numerous examples from opera, Rothstein concludes that
there is a national element involved: eighteenth-century approaches to bar line placement persist in French
and Italian operatic works well into the nineteenth century, long after they had been retired from German
musical practice.

In German metre, bars are e ectively containers for melodic phrases, and downbeats at the beginning of a C20.P35
C20.P36
piece coincide with the start of the melody, or at most, they start with a short anacrusis. In terms of Lerdahl
and Jackendo ’s (2010[1983]) theories of grouping and metre, they are ‘in phase’, or only slightly out of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
phase, in the case of a short anacrusis. By contrast, in Franco-Italian metre, grouping and metre are
frequently out of phase to a larger degree. Rothstein explains this as an e ect of the di erences between the
prosodic features of French and Italian verse, which are end-accented, and German, which is not, and in
which trochaic metre (a strong syllable followed by a weak one) is also common. In notational practice, this
leads to a di erent understanding of the role of the bar line, and prescriptions for its use. In the musical
p. 428 setting of Franco-Italian end-accented verse, and in common with earlier theories of notational practice,
the bar line should precede the resolution of the nal cadence, the location of the accento commune in Italian
23
verse, or accent tonique in French (i.e. the nal accent in a line of poetry). To put it crudely, to determine
where to put a bar line, the composer should look rst for where the nal stressed syllable of the line falls,
put a bar line in front of it, and work backwards. This may result in an ‘anacrusis’ of half a bar or more at the
beginning of the piece, which, according to Rothstein, is one of the characteristic features of a score barred
24
according the principles of Franco-Italian metre.

Interpreted in the light of the French song that it is based on, could the music in Figure 20.3 be an example C20.P37
of Franco-Italian hypermetre, as explained by Rothstein (2011)? Taking examples from Verdi’s mid-century
operas, Rothstein suggests that the presence of an uneven number of bars’ introduction to a melody, or an
apparently unnecessary half-bar rest in the rst bar, is an indicator of hypermetrical organization
combined with Franco-Italian barring rules. The prime example of this is Verdi’s aria ‘La donna è mobile’
from Rigoletto (1851), where the half-bar vamp at the beginning suggests a duple hypermetre, as a result of
which the textual stress falls, as seems more appropriate, on the rst syllable of mobile rather than la. In the
same way, ‘Monsieur Dumollet’ makes sense if viewed as hypermetrically in 12/8, the melody beginning on
the last beat. In this light, the dancers’ apparently unmusical three-beat vamp could be reread logically as
the rst three beats of a quadruple hypermeasure.

What remains unexplained, however, is the apparent illogicality or unmusicality of dancers counting ‘1’ and C20.P38
beginning a sequence on what appears now to be an even weaker beat, not with an anticipatory movement,
but with one of a series of identical steps. It is only illogical, however, if viewed within a framework of
‘German metrical hearing’, as Rothstein calls it (2011: 98), in which the rst beat is considered strong (1 2 3
4). Franco-Italian, end-accented metrical hearing can be characterized as leading in the other direction, 1 2
3 4, or expressed more clearly, 2 3 4 1. Roger Grant (2014: 213) also shows how these contrasting ‘front-
loaded’ and end-accented views of metre were represented in the nineteenth century by Gottfried Weber
and Jérôme-Joseph de Momigny respectively. Momigny suggested, for example, that musicians should
count phrases starting at 1 from the point at which they began, even when that was in the middle of the bar,
which shows a surprising agreement between the supposedly irreconcilable (at times) counting schemes of
musicians and dancers. His theory does not eliminate or change the concepts of upbeat and downbeat, but it
di erentiates between the sense of musical ‘proposition’, and where they begin and end, from the issue of
metre. From this perspective, it is a confusion of terms and concepts to regard the dancers’ ‘1’ as having
weight or accent, which would be a particularly German way of interpreting it metrically: as in Momigny’s
theory, it is the beginning of a proposition. In addition, to the extent that in a sequence which perhaps
moves from one side of the stage to the other from point A to point B, B is a destination in a similar sense to
the accento commune of a line of verse.

The next three examples are taken from the 2013 Grades 4–5 ballet examination syllabus of the RAD for C20.P39
p. 429 which I edited the piano scores. The problems I describe here illustrate the degree to which
representation in music can be a matter of localized decision making, contingent on speci c problems and
audiences.

Example 2: Bizetʼs LʼArlésienne C20.S6

Example 2 is taken from the Prelude to Bizet’s L’Arlésienne suite (1872). Figure 20.5 shows the position of C20.P40
the bar lines as they appear in the piano accompaniment for the exercise for grands battements in the RAD

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Grade 4 ballet examination syllabus (Royal Academy of Dance 2013: 78–79).

Figure 20.5

Bizet, Prelude from LʼArlésienne (1872), showing Bizetʼs original barring, and the placing of the dancersʼ ʻ1ʼ on the third beat of C20.F5
the bar.

This is Bizet’s barring, but the passage before the double bar line in the example was added as a four-count C20.P41
introduction. In RAD music publications, the double bar line is used near the beginning of a piece to indicate
where the introduction ends and the exercise proper begins. In discussion with the creators of the syllabus
and those charged with writing the syllabus (including its counts) in text form as well as Benesh Movement
Notation, I had to decide whether to re-bar the music in ‘German’ style or leave it as it was. In the end, I
could not bring myself to change the original, even though it would probably not make any di erence to the
way that listeners would hear it. Bizet’s barring seems ambiguous with regard to accent and makes room for
several interpretations, but the nal C of the melody (an instance of which is shown in bar 3 of Figure 20.5)
is always accented. In addition, the melody is a borrowing of a much earlier tune, ‘La marche des rois’,
usually sung as a Christmas carol. The rst line shows clearly the same principle of end-accentuation which
25
would explain Bizet’s barring: ‘[De] bon matin, J’ai rencontré le train’ (main accented syllables shown in
bold).

Moving Bizet’s bar lines might imply a German metre, where the third beat of the bar is weaker than the C20.P42
rst, and although there is no telling how individual pianists might interpret it, this seems wrong. There
was little choice but to write ‘1’ in brackets under the anacrusis in the piano score, to guard against well-
meaning pianists telling the teacher (in accordance with conventional principles of music notation) that the
rst count of the exercise was at the beginning of bar 3.

p. 430
Examples 3 and 4: Verdiʼs Otello C20.S7

The same problem of music that starts unexpectedly at the half bar arose with two extracts from Verdi’s C20.P43
ballet music for the Parisian première of Otello in 1894. First, the ‘Canzona greca’, used for the combined
exercise of ronds de jambe à terre and battements fondus in Grade 5 (Royal Academy of Dance 2013: 72–73).
Figure 20.6 is an expansion of the piano score with an extra line to show the woodwind line (marked ‘Ww.’)
26
that is played in the orchestral version, but omitted in the piano score.
Figure 20.6

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Verdi, ballet music from Otello (1894), ʻCanzona greca.ʼ C20.F6

The tonic harmony of the lower strings, and grace notes preceding the G# in the melody played by the C20.P44
violins and harp, establish an unmistakable metrical accent that contravenes any ‘Germanic’ interpretation
of the 4/4 metre. For this reason, as with the Bizet example in Figure 20.5, I retained Verdi’s barring, but
since the introduction began on the third beat of a bar, I added a ‘(1)’ under the rst note of it to clarify that
it should be regarded as count 1, rather than a half-bar anacrusis, and placed a double bar line in the middle
of bar 3 to show where the introductory four counts ended, and the exercise began.

While these additions help orientate users (i.e. pianists, dancers, notators, and dance teachers) to the C20.P45
necessary hearing of the music for the exercise, the music is still metrically ambiguous and multi-layered:
no single metrical level seems to prevail. The harmonic rhythm suggests an additional possibility of hearing
4/2 hypermeasures beginning on the half bar (shown with the brackets marked H1 in Figure 20.6), which
align with the dancers’ counts. Taking into account the high E of the woodwinds combined with the tonic
harmony of the rst beat of bar 3, and the similar event in bar 5 where a B in the woodwind is matched with
the dominant chord of B major, one might feel or place a greater metrical weight on the last beat than the
p. 431 rst in the H1 hypermeasures, supporting Rothstein’s view that one could count Franco-Italian metre 4 3
2 1, rather than 1 2 3 4. I have indicated this underneath the bottom stave, with crescendo marks to indicate
the direction of emphasis, rather than an actual increase in volume. On the other hand, the countermelody
in the woodwinds beginning in the second bar, especially with its instruction to be played ‘without accents’,
looks super cially as if it could also be hypermetrically in 4/2 starting half a bar later, and written according
to German principles—that is, framed by the bar lines or by the imaginary ones suggested by the
hypermeasures (marked H2).

The nal example is also from Verdi’s Otello: the Muranese used for ‘Turns’ in the RAD Grade 5 syllabus C20.P46
(Royal Academy of Dance, 2013: 102–103) (see Fig. 20.7). The numbers above the stave show the dancers’
counting scheme; ‘up up down’ indicates the movement pattern of the feet, which are brought up
alternately to knee height on 1 and 2, and stay down on 3 and 4.
Figure 20.7

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Verdi, ballet music from Otello, ʻMuraneseʼ. C20.F7

Without access to a score, the three-note slide to a forte, accented note at the beginning of the phrase, C20.P47
followed by an immediate piano, makes it hard not to hear the beginning of this melody as a downbeat, or a
‘1’ in dancers’ terms. At rst sight, this looks and sounds as if it might be a case of Franco-Italian metre, but
16 bars later the anacrusis disappears and a new theme starts on the rst beat of the bar. Has German metre
taken over at this midway point, or should one reconsider the status of the half-bar opening gure as a
straightforward anacrusis after all, not the beginning of an end-accented, out-of-phase melody? Either
way, the ambiguity of the metre at the beginning leads to complications in practice. To adapt the exercise to
the apparent metrical structure of the opening, as in the ‘Monsieur Dumollet’ example (Figures. 20.2–20.4),
by counting an introduction of three counts and then treating the anacrusis as ‘1’, would only work for half
27
the piece. The solution for the ballet exercise is to count the introduction as four, ignoring the heavy
accent on the fourth beat, and deliberately parsing it as an anacrusis. The turning steps then begin on the
downbeat of bar 3, which is likely to be perceived as weak in comparison to the loud anacrusis that precedes
it, even though it is strong in terms of the notated metre. At rst it might seem counterintuitive to count in
this way, but such a conceptualization of the metre is supported by the placement of cadences in the rest of
p. 432 the phrase and is essential if dancers are not to be thrown by a sudden shift in accent when a new melody
starts with a downbeat 16 bars later. With this in mind, it might be more appropriate to regard this passage
as exhibiting a form of ‘shadow metre’ (Samarotto 1999: 235, cited in London 2012: 101), i.e. a perceived
metrical structure that is at odds with the notated metre. What is particularly confusing here, though, is that
the shadow is heard before the (notated) metre that cast it.

These examples show that even in what would usually be regarded as simple Western music, counting is not C20.P48
as simple as it seems, or as it is often taught. Both of Rothstein’s publications end with a plea to accept that
there is no such thing as a single approach to metre, however universal its underlying principles may be.
Metre is a ‘culture bound phenomenon’ (Rothstein 2011: 110), and di erent approaches may be needed
according to the repertoire in question, particularly where poetry and singing are involved. He also points
out at the end of his discussion of Franco-Italian hypermetre (2011: 109) that national and historical
variations in the placement of bar lines require a rethinking, or localized nessing, of the concepts of metre
and accent. Under Franco-Italian metrical principles, barring helps to locate the chief linguistic stresses of a
line, so that ‘accent’ here means what Lerdahl and Jackendo (2010[1983]) would call phenomenal accent
(i.e. an exceptional increase in intensity in the sound in some way, as opposed to metrical or structural
accents). In German metre, accent refers to ‘ordinal primacy’ (Rothstein 2011: 109)—that is to say, it is
simply ‘ rst’ in the sense of its placement in the bar, rather than being necessarily perceptibly stronger. In
other words, it should be regarded as psychological rather than performed. If the rst beat of every bar is
played louder rather than merely thought of as salient by the performer, the results are ‘aesthetically
deplorable’ (p. 109). For this reason, Rothstein says, in the late nineteenth century, some German theorists
advocated a return to earlier, Franco-Italian metrical approaches, as a means of recovering the musical
sense and shape of phrases.
These issues become a matter of politics in that to teach metre, time signature, and the concept of metrical C20.P49
accent in accordance with ‘German’ metrical principles as part of a transhistorical, universal system, is an
ideological position that places those who do not hear metre in that way at a disadvantage. It also
contributes to myths of music and dance as universal languages, and risks misinterpreting music of the type
explored in this chapter. In the music-theoretical discipline in which Rothstein and others work, there are
28
competing views on and interpretations of metre and time signature which can be traced to individual
theorists and particular national musical and theoretical traditions, but this is rarely if ever acknowledged

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
in courses that claim to teach music theory to dancers.

Rothstein’s research is particularly helpful in understanding the nineteenth-century ballet repertoire, C20.P50
where there are many examples of Franco-Italian (hyper)metre. It helps to make sense of the relationships
between choreography and music where similar tensions between metre and grouping occur, and it allows
for a concept of metre that is internally contradictory and ambiguous. In the Tchaikovsky example (Figure
20.2, Example 1) above, for instance, the melody has a metrical structure of its own that is parallel to the
accompaniment, yet half a bar out of phase with it. These two metrical systems cannot be totally reconciled
p. 433 either theoretically or practically: within a notated metre of 6/8, they are e ectively in canon with each
other in a form of perpetual ‘shadow metre’. In this apparently simple music, something rather more
complex is happening.

If metre or time signature is taught abstractly as a neutral, universally applicable system without reference C20.P51
to particular works, these interesting, disruptive, and creative complexities disappear. In the context of
ballet teaching, a singular but misunderstood German-orientated interpretation of metre is in danger of
being a circular, self-ful lling prophecy, which is potentially worsened rather than improved by live music:
a pianist teaches the ballet teacher to count metrically, the teacher then constructs her exercises using this
model of metre and asks the pianist to improvise or perform music according to the pattern prescribed, and
so on. In turn, ballet teachers encourage children to listen to a piece of music and clap along in time, loudly
accenting the rst beat of each notional bar and perhaps counting at the same time. As the examples here
illustrate, if the aim is musical understanding, more might be gained by inviting divergent hearings than by
promoting a single ‘correct’ one.

The analytical methods developed by Stephanie Jordan provide one example of how things might be done C20.P52
di erently. These consist of combining elements of music and dance analysis, including notation, together
with ‘sketch learning’—trying out what it feels like to do movements along with the music. Describing such
a teaching episode with graduate students, Jordan refers to the way that this facilitated ‘the excitement of
direct experience of conversation with music, unmediated by counts’ (Jordan 2011: 56). What Jordan is
celebrating is not an idealized, direct, unmediated experience of dance and music (if, indeed, such a thing is
at all possible). Clearly, it is being mediated by her discursively in the act of teaching. Sketch learning,
perhaps with self-accompaniment through singing or listening to a recording, is a complement to, not a
replacement for, the score, counts, and verbal description. Jordan observes that her ideas ‘could have
implications for professional practice, renewal of the repertory, and the education of dancers and
choreographers, not simply for dance academics and students’ (2011: 61, n. 17). By focusing on
choreographic and musical works, together with their representations in notation, she is able to draw out
the tensions and contrasts between movement, music, and score, and develop an understanding of rhythm
as a complex, relational phenomenon, grounded in practice, rather than a universal rule-based system.

For musicians, an appreciation of the principles of Franco-Italian hypermetre can lead to new experiments C20.P53
in, and insights into, accentuation and phrasing. For ballet pianists in particular, it creates a feeling of what
teachers refer to as ‘up’. That exasperatingly vague request to the novice pianist, ‘Can we have something a
bit more up please?’ might not mean just ‘a bit faster’, but more precisely, something that is written and
played according to Franco-Italian barring principles. It may even be that one could talk about music with a
29
‘French’ or ‘Italian’ feel, and develop a sense for its metrical particularities that contradict conventional
metric counting.

Rothstein’s observations are illuminated further by Alison Stevens’s (2018) research on eighteenth-century C20.P54
contredanse music. The majority of French contredanse tunes were notated according to the Franco-Italian
principles that Rothstein describes, i.e. with a half-bar upbeat, whereas English ones were more likely to
p. 434 begin with a short anacrusis, or on a downbeat. This was not just a matter of notational rules, but of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
steps: unlike the English version, the French contredanse incorporated gavotte steps starting on the half-
bar, so that when tunes crossed the channel between France and England, bar line and downbeats moved
too, depending on the country (Stevens 2018: 42–47). As a consequence, ‘[u]sing Franco-Italian barring in
an opera, for example, might not only be linguistically appropriate but also feel French because of the dance
experience it evokes’ (p. 45). A knowledge of these texts and dances brings insights to the performance of
the music that studying the music alone does not. It unsettles not just the idea of music’s autonomy, but the
idea of what ‘knowledge of music’ means in the rst place, both for dancers and musicians. As Bohlman
says, the body ‘metaphysically loosens music from its own autonomy, mapping it on to other physical
practices, such as ritual and dance, which in Western terms have “nothing to do with music” ’ (Bohlman
2001: 33; see also Bohlman 1993: 434–435).

Conclusions C20.S8

As I have shown, it is problematic to assume that counting systems in music are universal and undisputed, C20.P55
or that counting is something to be left for musicians to do, while dancers get on with more gural, natural,
or authentic, embodied responses to music. Tensions between rhythm and metre are a topic for music
theorists themselves, despite the fact that in everyday usage, ‘music theory’ is often used to mean the
rudiments of notation, such as time signature. Christopher Hasty characterized the opposition in one pithy
sentence: ‘We can disparage a performance as being too metrical, but it would make no sense to say that a
performance is too rhythmic’ (Hasty 1997: 5). David Smyth (1992) reminds us that Edward Cone, credited
with inventing the term ‘hypermetre’, used it only relatively late in his work, and in a pejorative sense. The
greater burden of his argument was that rhythmic structure did not equate to large-scale metrical
organization: ‘I insist that on some level, this metric principle of parallel balance [i.e. the alternation of
strong and weak beats] must give way to a more organic rhythmic principle that supports the melodic and
harmonic shape of the phrase and justi es its acceptance as a formal unit (Cone 1968: 26, cited in Smyth
1992: 79).

Paradoxically, although the examples discussed here appear to address issues of notation, they are only C20.P56
issues at all because of what is not or cannot be represented: the culturally contingent notation and
versi cation practices; the metrical interpretations that di erent musicians will infer from the printed
score; listeners’ metrical parsing of music as heard; the teachers’ framing of the metre as they count or give
instruction over the music; the choreography performed to the music; the dancers’ perception of the music,
which they may have to ignore in order to bring whichever metrical frame best serves them to learn or
perform the choreography; and the possibility that, like the narrator of Swing Time, they might struggle to
make any sense of the combination of music, dance, and teachers’ counts at all. Without taking both the
p. 435 representational and non-representational into account, and holding them in tension, these issues—
which would seem important and interesting for interdisciplinary research in dance education, the training
of ballet teachers, and creative, innovative practice in music and dance—are in danger of disappearing from
view.
Acknowledgements C20.S9

My thanks to Douglas Corbin, Susie Cooper, Susie Crow, Christopher Hampson, Lucy Green, Wen Yang Ho, C20.P57
Stephanie Jordan, Helen Linkenbagh, Anna Wooster Pasti, and Vicki Watts for their comments and helpful
suggestions on early drafts of this chapter, and to Emily Payne for her editorial help throughout.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Notes
1. When I talk about the di erence between dancersʼ counts and metre, I mean instances where counts are mostly C20.N1
coordinated with beats, and where dancers have not completely misconstrued or ignored the metre, such as counting in
8s across the 6-beat structure of a polonaise, to take an extreme example. Movement is sometimes performed in silence,
or accompanied by music that has no clear perceptible beat pattern, but this is a di erent matter, and not my focus here
(but see Jordan 2018 for examples).

2. I selected this example because it is particularly ambiguous from a musicianʼs point of view, but the dozen or so dancers C20.N2
and coaches I asked about this passage all said that while they might count it when teaching someone else the
choreography, they would not do so themselves, because it seemed obvious how to place the steps on the music.
Nonetheless, whether the counting is done consciously or not, there is still a notional choreographic ʻ1ʼ on the 3rd beat of
the 2nd bar.

3. The counting of this passage seems bizarre, but less so when compared to the coda of the pas de deux (Act II No. 14) from C20.N3
Tachaikovskyʼs The Nutcracker (1892), which begins with the exactly the same 3-note motif in a di erent key, but in 2/4, so
that the 2nd note falls on the beginning of the bar, and is also given an accent by Tchaikovsky.

4. See also Cook (2018) for an extended discussion of Bambergerʼs ʻwipe-outʼ concept. C20.N4

5. Borrowed from Nattiezʼs terms poïetic and esthesic (Nattiez 1990). C20.N5

6. Although I was not aware of it until a er I had finished writing this chapter, Bell (2017: esp. 39–88) also contains a detailed C20.N6
music-theoretical discussion of Rothsteinʼs theory of national metrical types in relation to Tchaikovskyʼs ballet music.

7. A representative sample of the kind of choreomusicological studies I have in mind includes Damsholt (2002), Hodgins C20.N7
(1992), Jordan (2000; 2008; 2011; 2012; 2015; 2018), Mason (2012), and White (2006).

8. There are a few studies to my knowledge, such as Côte-Laurence (2000), Kerr-Berry (2001), and Pollatou et al. (2003), but C20.N8
not enough to suggest a serious interest in the topic.

9. A list of dance-awarding bodies validated by the Council for Dance Drama and Musical Theatre (formerly CDET) can be C20.N9
found at their website: https://cdmt.org.uk/validated-awarding-organisations.

p. 436 10. BMN has some similarities with music notation, such as the use of a 5-line stave, le -to-right reading direction, and the C20.N10
use of bar lines to show correspondences with music scores, where these exist, or have been used during the notating
process. BMN scores indicate the temporal alignment between choreography and score by the use of ʻpulse beatsʼ and
counts.

11. The same is true for any level of dance teaching, ballet particularly, but my focus in this chapter is on childrenʼs work C20.N11
because it a ects the largest number of teachers and students, and because the issues are regarded (wrongly, in my view)
as basic and rudimentary.

12. See also Barry et al. (2008); Born and Barry (2013). C20.N12

13. A humorous example: the ethnomusicologist Bruno Nettl has explained that while colleagues in his discipline have long C20.N13
considered dance to be part of their work, dance practitioners and academics in 20th-c. America have striven for
recognition as an independent art form: his daughter Rebecca Nettl-Fiol, dance professor and program co-ordinator at the
University of Illinois, ʻwouldnʼt be inclined to consider attaching her department to the School of Musicʼ, where her father
is Professor Emeritus of Musicology (Nettl 2010: 16).
14. There may be more accessible or suitable words in English than ʻobjectificationʼ, but since Faure uses objectivation in C20.N14
French, any other terms, such as ʻreductionʼ, ʻtranslationʼ, or ʻreificationʼ, seem to add nuances of meaning that are not in
the original text. I suspect that finding the right word is di icult, because the concept is complex. Justin London (1993)
characterized proponents of di erent views of metre as broadly either ʻstructuralistsʼ and ʻphenomenologistsʼ. According
to this scheme, structuralists regard music as an entirely external stimulus, and metre a kind of response to it—in other
words, metre is ʻout thereʼ in the musical sound. ʻPhenomenologistsʼ, by contrast, regard musical structure as the product
of an interaction between listener and sound, and metre as an activity, a counting-along-in-time that enables listeners to
make metrical predictions about the music, without needing every beat and metrical accent to be present in the sound.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
There are ontological implications for both metre and music here: metre is not part of the music, and ʻthe musicʼ, at least
for Londonʼs phenomenologists, includes something (the listenerʼs metrical engagement) that is not part of the musical
sound. From this point of view, counts are not simply extrapolations of the beat, but they also help constitute or construct
it. See London (2012: 9–24, 99–109) for more on metre as attention, and metrical ambiguity respectively.

15. ʻLʼobjectivation de la structuration musicale des mouvements dansés réduit les possibilités dʼinterprétation de la C20.N15
musiqueʼ.

16. Faure does not cite Bamberger directly, but refers to an outline of theory presented in Gardner (1993: 118). She cites the C20.N16
French translation of the book, I have given the English source in this chapter. See also Bamberger (1991; 2013).

17. Incidentally, one of the few scholarly studies of music and a dance-teaching society in the UK is by a cultural geographer, C20.N17
Tim Cresswell (2006). This seems to support Bornʼs idea that reconfiguring conceptual boundaries in musicology require
us to ʻlook outside, beyond the archipelago, to the key adjacent disciplines—the next-nearest knowledge continents–that
lie beyond musicology: that is, to the sciences of the cultural, social and temporalʼ (Born 2010: 210).

18. I am borrowing this concept from Goehr (2002: 14–15). C20.N18

19. In fact, since it is characteristic of NRT to develop ethnographic styles that account for the complexities of embodied C20.N19
experience (Vannini 2015), the evocation of musical experience in dance classes in Swing Time could be regarded in some
ways as a form of NRT in itself.

p. 437 20. Rather like the term ʻcommunityʼ which, as Elias says, is subject to the same romantic idealization. See also Nicolini (2013: C20.N20
94).

21. It is important to note, however, that DeNora (2003: 45) counts music analysts as a ʻsubset of music users and music C20.N21
consumersʼ.

22. See also Jacobsen (2009) on the same topic. C20.N22

23. See Baragwanath (2011: 66–89) for a detailed exposition of this aspect of Italian compositional theory. C20.N23

24. I use the term ʻanacrusisʼ guardedly, since it is usually assumed to mean the upbeat before a downbeat, the weak-beat C20.N24
preparation for the strong, whereas in these Franco-Italian examples, the anacrusis at the beginning is, so to speak, merely
a side e ect of the placement of the much later bar line before the accento commune, where the sense of the text dictates
its position.

25. In the carol, there is an additional 16th-note syllable at the beginning for the word ʻdeʼ. C20.N25

26. The syllabus music is recorded in two versions: one with orchestra and other ensembles, one with piano. C20.N26

27. It would also be unworkable in the context of this syllabus, where each exercise has two pieces of music assigned to it. The C20.N27
other one of the pair, a Russian folk dance tune ʻKarapetʼ, has a four-count introduction.

28. See e.g. Mirkaʼs (2014) expansion of Allanbrookʼs (1983) treatise on dance and metre, where she charts the change from C20.N28
thinking about time signature as a means of signifying tempo and a ect to a notational system which signified neither.

29. See also Patel (2008: 159–168) and VanHandel and Song (2010) on the subject of relations between rhythm and language C20.N29
in instrumental music.
References C20.S10

Allanbrook, W. J. (1983). Rhythmic gesture in Mozart: Le nozze di Figaro & Don Giovanni. University of Chicago Press. C20.P58
Google Scholar Google Preview WorldCat COPAC

Arendt, H. (1998). The human condition (2nd edn). University of Chicago Press. C20.P59
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Bamberger, J. (1991). The mind behind the musical ear. Harvard University Press. C20.P60
Google Scholar Google Preview WorldCat COPAC

Bamberger, J. (2013). Discovering the musical mind: A view of creativity as learning. Oxford University Press. C20.P61
Google Scholar Google Preview WorldCat COPAC

Baragwanath, N. (2011). The Italian traditions and Puccini: Compositional theory and practice in nineteenth-century opera. Indiana C20.P62
University Press.
Google Scholar Google Preview WorldCat COPAC

Barry, A., Born, G., and Weszkalnys, G. (2008). Logics of interdisciplinarity. Economy and Society 37(1): 20–49. C20.P63
https://doi.org/10.1080/03085140701760841
Google Scholar WorldCat

Becker, J. O. (2004). Deep listeners: Music, emotion, and trancing. Indiana University Press. C20.P64
Google Scholar Google Preview WorldCat COPAC

Bell, M. T. (2017). Rhythmic gesture in classic ballet: Awakening Tchaikovskyʼs Sleeping Beauty. Doctoral dissertation, University of C20.P65
Texas. https://doi.org/10.15781/T2WP9TD8M
Google Scholar Google Preview WorldCat COPAC

Biddle, I. D. (2011). Listening, consciousness, and the charm of the universal: What it feels like for a Lacanian. In D. Clarke and C20.P66
E. Clarke (eds), Music and consciousness: Philosophical, psychological and cultural perspectives, 65–77. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Bohlman: V. (1993). Musicology as a political act. Journal of Musicology 11(4): 411–436. https://doi.org/10.2307/764020 C20.P67
Google Scholar WorldCat

Bohlman: V. (2001). Ontologies of music. In N. Cook and M. Everist (eds), Rethinking music, 17–34. Oxford University Press. C20.P68
Google Scholar Google Preview WorldCat COPAC

p. 438 Born, G. (2010). For a relational musicology: Music and interdisciplinarity, beyond the practice turn. Journal of the Royal Musical C20.P69
Association 135(2): 205–243. https://doi.org/10.1080/02690403.2010.506265
Google Scholar WorldCat

Born, G., and Barry, A. (2013). Interdisciplinarity: Reconfigurations of the social and natural sciences. In Interdisciplinarity: C20.P70
Reconfigurations of the social and natural sciences, 1–56. Routledge.
Google Scholar Google Preview WorldCat COPAC

Bull, M. (2000). Sounding out the city: Personal stereos and the management of everyday life. Berg. C20.P71
Google Scholar Google Preview WorldCat COPAC

Bull, M. (2007). Sound moves: iPod culture and urban experience. Routledge. C20.P72
Google Scholar Google Preview WorldCat COPAC

Bull, M., and Back, L. (eds) (2015). The auditory culture reader. Bloomsbury Academic. C20.P73
Google Scholar Google Preview WorldCat COPAC

Capelle: (1811). La clé du caveau, à lʼusage de tous les chansonniers français, des amateurs, auteurs, acteurs du vaudeville and de C20.P74
tous les amis de la chanson. Capelle & Renard, de lʼImpr. de Richomme. https://gallica.bnf.fr/ark:/12148/bpt6k11753517
Google Scholar Google Preview WorldCat COPAC

Clarke, E. (2005). Ways of listening: An ecological approach to the perception of musical meaning. Oxford University Press. C20.P75
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Clarke, E., Dibben, N., and Pitts, S. (2010). Music and mind in everyday life. Oxford University Press. C20.P76
Google Scholar Google Preview WorldCat COPAC

Cone, E. T. (1968). Musical form and musical performance. Norton. C20.P77


Google Scholar Google Preview WorldCat COPAC

Cook, N. (1992). Music, imagination, and culture. Oxford University Press. C20.P78
Google Scholar Google Preview WorldCat COPAC

Cook, N. (2018). Getting out of the garret: Social and material dimensions of music as creative practice. In N. Donin (ed.), The C20.P79
Oxford handbook of the creative process in music. Oxford University Press.
http://www.oxfordhandbooks.com/view/10.1093/oxfordhb/9780190636197.001.0001/oxfordhb-9780190636197-e-16
Google Scholar Google Preview WorldCat COPAC

Côté-Laurence: (2000). The role of rhythm in ballet training. Research in Dance Education 1(2): 173–191. C20.P80
https://doi.org/10.1080/713694263
Google Scholar WorldCat

Cresswell, T. (2006). ʻYou cannot shake that shimmy hereʼ: Producing mobility on the dance floor. In On the move: Mobility in the C20.P81
modern Western world, 123–145. Routledge.
Google Scholar Google Preview WorldCat COPAC

Damsholt, I. (2002). The marriage of music and dance: Understanding the world of choreomusical relations through a gendered C20.P82
metaphor. In M. Dithmer (ed.), Of another world: Dancing between dream and reality, 237–250. Museum Tusculanum Press.
Google Scholar Google Preview WorldCat COPAC

Davies, S. (2003). Themes in the philosophy of music. Oxford University Press. C20.P83
Google Scholar Google Preview WorldCat COPAC

DeNora, T. (2000). Music in everyday life. Cambridge University Press. C20.P84


Google Scholar Google Preview WorldCat COPAC

DeNora, T. (2003). A er Adorno: Rethinking music sociology. Cambridge University Press. C20.P85
Google Scholar Google Preview WorldCat COPAC

DeNora, T. (2014). Making sense of reality: Culture and perception in everyday life. Sage. C20.P86
Google Scholar Google Preview WorldCat COPAC

Dewsbury, J. D. (2011). Dancing: The secret slowness of the fast. In T. Creswell and P. Merriman (eds), Geographies of mobility: C20.P87
Practices, spaces, subjects, 51–67. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Dowling: (1995). Discipline and mathematise: The myth of relevance in education. Perspectives in Education 16(2): 209–226. C20.P88
Google Scholar WorldCat

Dowling: (1998). The sociology of mathematics education: Mathematical myths, pedagogic texts. Falmer Press. C20.P89
Google Scholar Google Preview WorldCat COPAC

Elias, N. (1998[1978]). On the concept of everyday life. In N. Elias, J. Goudsblom, and S. Mennell (eds), The Norbert Elias reader: A C20.P90
biographical selection, 166–174. Blackwell.
Google Scholar Google Preview WorldCat COPAC

Faure, S. (2000). Apprendre par corps: Socio-anthropologie des techniques de danse. La Dispute. C20.P91
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Gardner, H. (1993). Frames of mind: The theory of multiple intelligences (2nd edn). Fontana. C20.P92
Google Scholar Google Preview WorldCat COPAC

Goehr, L. (1992). The imaginary museum of musical works. Clarendon Press. C20.P93
Google Scholar Google Preview WorldCat COPAC

Goehr, L. (2002). The quest for voice: On music, politics, and the limits of philosophy. Oxford University Press. C20.P94
Google Scholar Google Preview WorldCat COPAC

p. 439 Gopinath, S. S., and Stanyek, J. (eds) (2014). The Oxford handbook of mobile music studies, vol. 1. Oxford University Press. C20.P95
Google Scholar Google Preview WorldCat COPAC

Gopinath, S. S., and Stanyek, J. (eds) (2017). The Oxford handbook of mobile music studies, vol. 2. Oxford University Press. C20.P96
Google Scholar Google Preview WorldCat COPAC

Grant, R. M. (2014). Beating time and measuring music in the early modern era. Oxford University Press. C20.P97
Google Scholar Google Preview WorldCat COPAC

Green, L. (2008). Music on deaf ears: Musical meaning, ideology and education (2nd edn). Arima. C20.P98
Google Scholar Google Preview WorldCat COPAC

Gritten, A. (2011). Distraction in polyphonic gesture. In A. Gritten and E. King (eds), New perspectives on music and gesture, 99– C20.P99
122. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Hasty, C. F. (1997). Meter as rhythm. Oxford University Press. C20.P100


Google Scholar Google Preview WorldCat COPAC

Herbert, R. (2011). Everyday music listening: Absorption, dissociation and trancing. Routledge. C20.P101
Google Scholar Google Preview WorldCat COPAC

Hodgins, P. (1992). Relationships between score and choreography in twentieth-century dance: Music, movement and metaphor. C20.P102
Edwin Mellen.
Google Scholar Google Preview WorldCat COPAC

Howe Jr, E. (1843). The Complete Preceptor for the Accordeon, containing a scale for the common or whole toned, and also a scale C20.P103
for the semitoned or perfect accordeon; together with a large collection of popular and fashionable music, arranged expressly for
this instrument. Elias Howe. https://imslp.org/wiki/The_Complete_Preceptor_for_the_Accordeon_(Howe%2C_Elias)
Google Scholar Google Preview WorldCat COPAC

Hynd, R. (1961). Dancing to music. Musical Times 102(1415): 28. https://doi.org/10.2307/948679 C20.P104
Google Scholar WorldCat

Jacobsen, M. H. (2009). Introduction: The everyday—An introduction to an introduction. In M. H. Jacobsen (ed.), Encountering C20.P105
the everyday, 1–41. Palgrave Macmillan.
Google Scholar Google Preview WorldCat COPAC
Jordan, S. (2000). Moving music: Dialogues with music in twentieth-century ballet. Dance Books. C20.P106
Google Scholar Google Preview WorldCat COPAC

Jordan, S. (2008). Stravinsky dances: Re-visions across a century. Dance Books. C20.P107
Google Scholar Google Preview WorldCat COPAC

Jordan, S. (2011). Choreomusical conversations: Facing a double challenge. Dance Research Journal 43(1): 43–64. C20.P108
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Jordan, S. (2012). Moving ʻchoreomusicallyʼ: Between theory and practice. Cahiers de la Société québécoise de recherche en C20.P109
musique 13(1–2): 11. https://doi.org/10.7202/1012345ar
Google Scholar WorldCat

Jordan, S. (2015). Mark Morris: Musician, choreographer. Dance Books. C20.P110


Google Scholar Google Preview WorldCat COPAC

Jordan, S. (2018). Choreographers and musicians in collaboration, from the twentieth to the twenty-first century. In N. Donin C20.P111
(ed.), The Oxford handbook of the creative process in music. https://doi.org/10.1093/oxfordhb/9780190636197.013.6
Google Scholar Google Preview WorldCat COPAC

Kassabian, A. (2013a). Ubiquitous listening: A ect, attention, and distributed subjectivity. University of California Press. C20.P112
Google Scholar Google Preview WorldCat COPAC

Kassabian, A. (2013b). You say invisible, I say ubiquitous: A (formally former) studentʼs response to Philip Taggʼs ʻCaught on the C20.P113
back foot: Epistemic inertia and visible musicʼ. IASPM@Journal 3(2): 86–95.
Google Scholar WorldCat

Kerr-Berry, J. A. (2001). Applications of Dalcroze Eurhythmics to dance training: Grasping mixed meter. Journal of Dance C20.P114
Education 1(3): 106–114. https://doi.org/10.1080/15290824.2001.10387188
Google Scholar WorldCat

Kossen-Veenhuis, T. H. M. F. (2017). ʻThe great thing about collaboration is that it never is perfectʼ: An ethnography of music and C20.P115
dance collaborations in progress. Doctoral dissertation, University of Edinburgh. http://hdl.handle.net/1842/25685
Google Scholar Google Preview WorldCat COPAC

Lanza, J. (1995). Elevator music: A surreal history of muzak, easy-listening and other moodsong. Quartet. C20.P116
Google Scholar Google Preview WorldCat COPAC

Lefebvre, H. (2004). Rhythmanalysis: Space, time, and everyday life (trans. S. Elden and G. Moore). Continuum. C20.P117
Google Scholar Google Preview WorldCat COPAC

p. 440 Lerdahl, F., and Jackendo , R. (2010[1983]). A generative theory of tonal music. MIT Press. C20.P118
Google Scholar Google Preview WorldCat COPAC

London, J. (1993). Loud rests and other strange metric phenomena (or, Meter as heard). Music Theory Online 0(2). C20.P119
http://www.mtosmt.org/issues/mto.93.0.2/mto.93.0.2.london.art.html
Google Scholar WorldCat

London, J. (2012). Hearing in time: Psychological aspects of musical meter (2nd edn). Oxford University Press. C20.P120
Google Scholar Google Preview WorldCat COPAC

Lorimer, H. (2005). Cultural geography: The busyness of being `more-than-representationalʼ. Progress in Human Geography C20.P121
29(1): 83–94. https://doi.org/10.1191/0309132505ph531pr
Google Scholar WorldCat
Lorimer, H. (2007). Cultural geography: Worldly shapes, di erently arranged. Progress in Human Geography 31(1): 89–100. C20.P122
https://doi.org/10.1177/0309132507073540
Google Scholar WorldCat

Lorimer, H. (2008). Cultural geography: Non-representational conditions and concerns. Progress in Human Geography 32(4): C20.P123
551–559. https://doi.org/10.1177/0309132507086882
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Mason, H. (2012). Music, dance and the total art work: Choreomusicology in theory and practice. Research in Dance Education C20.P124
13(1): 1–20. https://doi.org/10.1080/14647893.2011.651116
Google Scholar WorldCat

McCormack, D. P. (2002). A paper with an interest in rhythm. Geoforum 33(4): 469–485. https://doi.org/10.1016/S0016- C20.P125
7185(02)00031-3
Google Scholar WorldCat

McCormack, D. P. (2008). Geographies for moving bodies: Thinking, dancing, spaces. Geography Compass 2(6): 1822–1836. C20.P126
https://doi.org/10.1111/j.1749-8198.2008.00159.x
Google Scholar WorldCat

McCormack, D. P. (2013). Refrains for moving bodies: Experience and experiment in a ective spaces. Duke University Press. C20.P127
Google Scholar Google Preview WorldCat COPAC

Mirka, D. (2014). Topics and meter. In M. Danuta (ed.), The Oxford handbook of topic theory, 357–380. Oxford University Press. C20.P128
Google Scholar Google Preview WorldCat COPAC

Nash, C. (2000). Performativity in practice: Some recent work in cultural geography. Progress in Human Geography 24(4): 653– C20.P129
664. https://doi.org/10.1191/030913200701540654
Google Scholar WorldCat

Nattiez, J.-J. (1990). Music and discourse: Toward a semiology of music (trans. C. Abbate). Princeton University Press. C20.P130
Google Scholar Google Preview WorldCat COPAC

Nettl, B. (2010). Nettlʼs elephant: On the history of ethnomusicology. University of Illinois Press. C20.P131
Google Scholar Google Preview WorldCat COPAC

Nicolini, D. (2013). Practice theory, work, and organization: An introduction. Oxford University Press. C20.P132
Google Scholar Google Preview WorldCat COPAC

Patel, A. D. (2008). Music, language, and the brain. Oxford University Press. C20.P133
Google Scholar Google Preview WorldCat COPAC

Pollatou, E., Hatzitaki, V., and Karadimou, K. (2003). Rhythm or music? Contrasting two types of auditory stimuli in the C20.P134
performance of a dancing routine. Perceptual and Motor Skills 97: 99–106.
Google Scholar WorldCat

Pollner, K., and Emerson, R. M. (2001). Ethnomethodology and ethnography. In P. Atkinson, A. Co ey, S. Delamont, J. Lofland, C20.P135
and L. Lofland (eds), Handbook of ethnography, 118–135. Sage.
Google Scholar Google Preview WorldCat COPAC

Pouillaude, F. (2017). Unworking choreography: The notion of the work in dance (trans. A. Pakes). Oxford University Press. C20.P136
Google Scholar Google Preview WorldCat COPAC

Quinones, M. G., Kassabian, A., and Boschi, E. (eds) (2013). Ubiquitous musics: The everyday sounds that we donʼt always notice. C20.P137
Ashgate.
Google Scholar Google Preview WorldCat COPAC

Revill, G. (2004). Cultural geographies in practice. Performing French folk music: Dance, authenticity and nonrepresentational C20.P138
theory. Cultural Geographies 11(2): 199–209. https://doi.org/10.1191/14744744004eu302xx
Google Scholar WorldCat

Rothstein, W. (2008). National metrical types in music of the eighteenth and early nineteenth centuries. In D. Mirka and K. Agawu C20.P139
(eds), Communication in eighteenth-century music, 112–159. Cambridge University Press.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Rothstein, W. (2011). Metrical theory and Verdiʼs midcentury operas. Dutch Journal of Music Theory 16(2): 93–111. C20.P140
Google Scholar WorldCat

p. 441 Royal Academy of Dance (2013). Graded examinations in dance. Ballet grades 4–5: Male/female music for set exercises and dances. C20.P141
Royal Academy of Dance.
Google Scholar Google Preview WorldCat COPAC

Samarotto, F. (1999). Strange dimensions: Regularity and irregularity in deep levels of rhythmic reductions. In C. Schachter and C20.P142
H. Siegel (eds), Schenker Studies 2, 222–238. Cambridge University Press.
Google Scholar Google Preview WorldCat COPAC

Schwartz, H. (2011). Making noise: From Babel to the big bang and beyond. Zone Books. C20.P143
Google Scholar Google Preview WorldCat COPAC

Shepherd, J., Virden, P., Vulliamy, G., and Wishart, T. (1977). Whose music? A sociology of musical languages. Latimer. C20.P144
Google Scholar Google Preview WorldCat COPAC

Smith, Z. (2017). Swing time. Penguin. C20.P145


Google Scholar Google Preview WorldCat COPAC

Smyth, D. (1992). Patterning beyond hypermeter. College Music Symposium 32: 79–98. C20.P146
Google Scholar WorldCat

Sterne, J. (2003). The audible past: Cultural origins of sound reproduction. Duke University Press. C20.P147
Google Scholar Google Preview WorldCat COPAC

Sterne, J. (2012). MP3: The meaning of a format. Duke University Press. C20.P148
Google Scholar Google Preview WorldCat COPAC

Stevens, A. (2018). Motion as music: Hypermetrical schemas in eighteenth-century contredanses. Masterʼs thesis, University of C20.P149
Massachusetts. https://scholarworks.umass.edu/masters_theses_2/703
Google Scholar Google Preview WorldCat COPAC

Still, J. (2015). How down is a downbeat? Feeling meter and gravity in music and dance. Empirical Musicology Review 10(1–2): C20.P150
121–134.
Google Scholar WorldCat

Stockfelt, O. (1997). Adequate modes of listening. In D. Schwarz et al. (eds), Keeping score: Music, disciplinarity, culture, 129–146. C20.P151
University Press of Virginia.
Google Scholar Google Preview WorldCat COPAC

Sulcas, R. (2013, March 27). Taking ballet to new heights. New York Times. C20.P152
https://www.nytimes.com/2013/03/28/arts/artsspecial/6-ballet-directors-discuss-current-evolution-of-form.html
WorldCat
Tagg, P. (2011). Caught on the back foot: Epistemic inertia and visible music. Keynote presentation, IASPM 16th Biennial C20.P153
Conference.
Google Scholar Google Preview WorldCat COPAC

Tagg, P. (2012). Musicʼs meanings: A modern musicology for non-musos. Music Scholarsʼ Press. C20.P154
Google Scholar Google Preview WorldCat COPAC

Taylor, V. (2003). Ballerinas in the church hall: Ideologies of femininity, ballet, and dancing schools. Doctoral dissertation, C20.P155

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353470863 by National Science & Technology Library user on 26 May 2023
University of Southampton. http://eprints.chi.ac.uk/1612/
Google Scholar Google Preview WorldCat COPAC

Thomson, V. (1962[1939]). The state of music. Vintage Books. C20.P156


Google Scholar Google Preview WorldCat COPAC

Thri , N. (1997). The still point: Resistance, expressive embodiment and dance. In S. Pile and M. Keith (eds), Geographies of C20.P157
resistance, 124–151. Routledge.
Google Scholar Google Preview WorldCat COPAC

VanHandel, L., and Song, T. (2010). The role of meter in compositional style in nineteenth century French and German art song. C20.P158
Journal of New Music Research 39(1): 1–11. https://doi.org/10.1080/09298211003642498
Google Scholar WorldCat

Vannini, P. (2015). Non-representational ethnography: New ways of animating lifeworlds. Cultural Geographies 22(2): 317–327. C20.P159
https://doi.org/10.1177/1474474014555657
WorldCat

Warren, G. W. (1989). Classical ballet technique. University of South Florida Press. C20.P160
Google Scholar Google Preview WorldCat COPAC

White, B. (2006). ʻAs if they didnʼt hear the musicʼ or: How I learned to stop worrying and love Mickey Mouse. Opera Quarterly C20.P161
p. 442 22(1): 65–89.
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
CHAPTER

21 Temporalities of North Indian ClassicalC21Listening: How


Listeners Use Music to Construct Time 
Chloё Alaghband-Zadeh

https://doi.org/10.1093/oxfordhb/9780190947279.013.23 Pages 445–466


Published: 08 December 2021

Abstract
This chapter explores how ethnography with musical listeners can illuminate relationships between
music and time. While much existing scholarship equates musical temporalities with qualities of the
‘music itself’, this chapter addresses the need for research that considers the diverse ways listeners
use music to engender experiences of time. Alaghband-Zadeh focuses here on rasikas, connoisseurs of
North Indian classical music. She shows how rasikas construct and experience North Indian classical
performances as sites of leisurely temporality: this is both an ethical practice, aligned with ideas of
virtue, and also a means for rasikas to position themselves as set apart from a world they view as
increasingly characterized by speed. Alaghband-Zadeh argues that music is a powerful temporal
resource: a means through which people cultivate ways of inhabiting time. Moreover, the immediate
temporalities of live performances contribute to the production of broader, public temporalities of
modernity, changing social formations and imagined histories.

Keywords: listening, listeners, North Indian classical music, rasikas, India, slowness, acceleration, ethics,
temporal resource, ethnography
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

WHAT is the relationship between the moment-to-moment temporalities of musical performance and the C21.P1
broader temporalities of history, people’s world-views, society, nation and modernity? In this chapter, I
address this question through an ethnographic study of ways of listening to North Indian classical music. I
suggest that existing literature on music and time has not paid enough attention to the particular ways in
which music is put to use by di erent constituencies of listeners, nor to how musical temporalities are
shaped by the material, spatial, and embodied contexts of musical performance. As a result, it has tended to
overemphasize structural aspects of the musical sound as the sole determinants of musical temporalities.
Moreover, by failing to account for the important role listeners’ aesthetic experience plays in mediating
between musical temporalities and temporalities beyond music, this work has been unable to address ways
in which multiple, intersecting temporalities come together in the moment of performance.

By contrast, I take an approach that centres listeners’ aesthetic experience; I do this through ethnography C21.P2
and interviews within a particular contemporary culture of listening based in India. I argue that
ethnographic work with listeners can reveal crucial insights into how music constructs time. My research
into this South Asian musical scene suggests that music functions as a powerful temporal resource: listeners

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
(and musicians) use music to cultivate particular ways of inhabiting time. In the North Indian classical
musical context, I show how a particular group of elite listeners constructs live performances as sites of
p. 446 leisurely temporality, removed from the hustle and bustle of everyday life and evocative of an imagined,
more leisurely time in the past. This way of constructing time through music is conditioned not only by the
a ordances of the musical sound (including aspects of musical structure) but also by the materialities of the
performance context (its spaces, technologies, and patterns of embodiment). Musical temporalities are also
enmeshed in broader social processes that extend far beyond the performance. As I demonstrate, the
leisurely temporality that these listeners construct and experience in North Indian classical music is central
to how they position themselves within historical time and in relation to changing social formations in
contemporary India. I therefore argue that musical performance is a site that brings together temporalities
across multiple scales: the immediate temporalities of the performance, co-produced by musicians and
listeners, contribute to the production of broader, public temporalities of modernity, changing social
formations and imagined histories.

Putting Musical Temporalities in Context C21.S1

A key aim for the academic literature on music and time (from across the musical disciplines) has been to C21.P3
explore the ways in which musical temporalities are tied to other, often larger-scale temporalities that
operate in the lives of music’s performers, composers, and listeners. For example, in ethnomusicology, this
impulse underscored Judith Becker’s (1979) seminal work on the temporalities of Javanese gamelan music,
in which she highlighted powerful parallels between the way time is organized in musical performance and
the organization of time within Javanese society more broadly, including at the scale of calendrical time.
Likewise in musicology, Karol Berger has famously argued that changes in the temporal organization of
Western classical music between the time of Bach and Mozart mirrored broad transformations in how
people in Europe conceived of time in general; he characterized this as a move towards thinking of time as
linear, both in music and beyond, re ecting a shift ‘from the pre-modern Christian moral-political outlook
to the modern post-Christian worldview’ (2008: 16). This work has highlighted the importance of situating
musical constructions of time within broader historical, social, and cultural contexts, echoing the wider
concerns of ethnomusicology, popular music studies, and the so-called ‘new’ musicology. Indeed, Susan
McClary considered ‘the ability of music to produce constructions of temporality’ to be ‘one of its primary
means of performing cultural work’ (2000: 161); speci cally, she writes, temporality is a ‘case in point’ for
understanding how music shapes subjectivities and therefore contributes to broad cultural shifts.

However, despite this commitment to understanding musical time in context, there has been surprisingly C21.P4
1
little ethnographic scholarship on how people experience time when they listen to music. Rather, work in
this eld has tended to rely on musical analysis in order to characterize the temporalities that music
p. 447 produces, often equating musical time with structural features of the music (see also Hyland 2016;
Johnson 2015; Murphy 2018; Rowell 1979; Taylor 2016). While this body of literature has shed important
light on how di erent modes of musical organization can condition people’s experiences of time, I suggest
that sound-based or score-based understandings o er only a partial picture of musical temporality; by
locating musical time in the ‘music itself’, this work attens out the particularities of how di erent
listeners (and di erent constituencies of listeners) inhabit time in the moment of musical performance,
collapsing potentially diverse individual and collective experiences of time in music into speci c
characteristics of the musical sound. This also fails to account for the ways people’s experiences of musical
time are shaped by the material, spatial, and embodied contexts in which music is performed and heard.

Likewise, listeners’ experiences also have tended to be absent from the ways in which scholars have C21.P5
accounted for possible relationships between the temporalities of musical performance and the
construction of time beyond music contexts (e.g. the temporalities of a society or epoch). Having located

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
musical time in the musical sound or score, work in this vein (including the examples by Becker and Berger
already cited) then typically draws a link between musical and wider temporalities by pointing out
similarities in the ways time is organized in each of these domains. This is a classic structuralist move:
structural aspects of the music resemble other constructions of time in the culture from which that music
emerged, suggesting a relationship between them. However, as various scholars have pointed out, merely
highlighting similarities between musical time and time beyond music does not account for the possible
mechanisms through which they might be linked (Bar-Yosef 2001: 423–426; Clayton 2000: 6–7; McGraw
2008: 140–142); in particular, this risks neglecting the crucial role that listeners’ experiences play in
mediating between the a ordances of the musical sound on the one hand and, on the other, the production
of large-scale temporalities of imagined histories, world-views, nation, and modernity. Moreover, this
approach risks oversimplifying relationships between musical time and time beyond music: by
understanding these relationships in terms of resemblances between di erent kinds of structural
organization, this work emphasizes relationships that are grounded in similarity, foreclosing the possibility
of other, more complex kinds of relationships, such as those involving disjuncture or di erence.

By contrast, Georgina Born has argued that the temporality of the musical score or sound is only one level at C21.P6
which music shapes time. She stresses ‘the need to analyze the multiplicity of time in cultural production
(including music)’ (2015: 362), suggesting that musical objects engender temporalities on at least four
di erent levels. Rather than understanding musical time as determined solely by qualities of the musical
sound, she characterizes music as animating multiple, intersecting temporalities at di erent scales. Born’s
insistence on the multiplicities of musical time echoes recent work in media studies. There, Emily Keightley
has likewise highlighted ‘the multiple and various temporalities which are supported, shaped and
performed through mediated cultural life’ (2013: 55–56). Keightley argues that temporalities of late
modernity cannot be understood through analysing media texts or technologies alone, but rather that this
p. 448 also requires an examination of how people put these texts and technologies to use in their everyday
lives. She understands mediated time as produced in what she labels ‘zones of intermediacy’, drawing on
and adapting Henri Bergson’s concept of ‘zones of indeterminacy’ (1911): she uses this idea to theorize
experiences of time as emerging out of particular milieux, in which ‘time is multiply mediated by […]
technologies, texts and social contexts in situ’ (Keightley 2013: 71). In contrast to Bergson, Keightley
stresses the role of the experiencing subject who must reconcile the competing temporal logics that
intersect in any such milieu (see especially pp. 66–67); she notes that this analysis highlights the power
relations that shape mediated experiences of time, since the ways people navigate intersecting
temporalities are dependent on their di erent ‘competencies, resources and structural positioning’ (p. 68).
This work issues a powerful invitation to music studies to look beyond the musical sound in order to
understand how music shapes time, instead considering the multiple, intersecting (and potentially
competing) temporalities that come together in speci c acts of performing and listening to music, as well as
the ways in which people make use of the temporal possibilities that music a ords.

In this chapter, I suggest that relations between musical time and time beyond music can be brought into C21.P7
better focus through ethnographic work that centres listeners’ experiences of time, and that attends to the
material, embodied nature of the temporalities that musical performances animate. I present data from a
larger ethnographic study of musicians and listeners of North Indian classical music in three North Indian
cities (Delhi, Mumbai, and Pune), including participant observation and interviews with listeners and
2
performers. I also draw on my own experiences of listening to and learning to sing North Indian classical
3
music since 2004. Through this ethnographic work, I explore how listeners inhabit time in the moment of
musical performance; in doing so, I highlight a particular set of discursive and embodied practices, highly
characteristic of North Indian classical performances, through which certain elite listeners use this music to
construct and situate themselves in time. As I will show, taking this ethnographic approach makes it
possible to illuminate the ways in which the temporalities of musical experience emerge out of dynamic
encounters between musical sound and listeners, within (and shaped by) particular material, embodied,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
historical, and social contexts.

The listeners I focus on here refer to themselves as rasikas (connoisseurs). They are expert music-lovers and C21.P8
they form an important segment of the audience for North Indian classical music. Many possess great
musical expertise; some are amateur musicians. They occupy a high social status both in the music world
and beyond it. Within the North Indian classical music scene, being a rasika is a valued social identity;
listeners who are known for their musical knowledge and enthusiasm earn respect both from fellow
listeners and from musicians. Moreover, as I have argued elsewhere, rasikas embody a middle-class habitus,
4
reproducing values of what in India is an elite social group (despite the misnomer ‘middle class’). At live
performances, rasikas typically try to sit as close as possible to the musicians, and show their appreciation
for the music by gesturing or commenting out loud throughout the performance. These listening practices
p. 449 constitute what Deborah Kapchan calls a ‘genre of listening’ (2015): they involve a very particular set of
embodied and discursive practices, which in turn are linked to wider musical ideologies that circulate within
5
this musical scene.

Although I had not initially intended my research with rasikas to explore questions of time and temporality, C21.P9
time frequently came up as a topic in conversation when I interviewed them about their listening
experiences. Through these conversations, it became evident that for rasikas one of the key listening
pleasures this music a ords is a particular way of experiencing time: they credit North Indian classical
music with making possible the experience of a kind of slowed-down, leisurely temporality, which they
understand as set apart from the speeded-up temporality of contemporary lives. As I will show in this
chapter, the leisurely temporality that rasikas experience in live performances also contributes to the
production of temporalities at much longer scales: this is central to the ways rasikas situate themselves in
historical time, in the context of contemporary India. Moreover, these temporal practices are part of how
these listeners construct themselves as ethical subjects. For rasikas, certain ways of experiencing time in
performance are infused with a sense of virtue and are tied to ideas about what it means to be a good
listener; I situate these attitudes within a wider, middle-class moral panic about Indian modernity, which
rasikas understand in terms of changing patterns of consumption and, importantly, changing attitudes to
time. Overall, I argue that music is what Daniel Paiva, Herculano Cachinho, and Teresa Barata-Salgueiro
(2017) call a temporal resource. It is a potent resource for listeners (and musicians) to cultivate embodied,
a ective experiences of time in the moment of performance or listening. Looking beyond the immediate
time of the performance, these experiences in turn contribute to the production of temporalities of nation,
modernity, and imagined histories: I suggest that the immediate temporalities of performance are a means
through which these larger-scale temporalities are made real in people’s everyday lives.

A Space of Leisurely Temporality C21.S2

When I interviewed rasikas about their listening experiences, the topic of time came up often. For example, C21.P10
Yavit, a prominent music patron, told me:

I always say when I am listening, the most important thing is: take your watches o ! Suspend that C21.P11
concept of time that you have! (OK, I will listen one hour forty minutes.) Let the music ow! Let it
ow through you!

With these remarks, Yavit describes the practice of listening to North Indian classical music as a practice C21.P12
that has its own distinct temporality. For him, being able to experience this music in the best possible way
relies on cultivating a temporal sensibility that is distinct from the customary ‘concept of time’ that people
have; it involves taking a relaxed attitude to the passage of time, taking one’s watch o and not setting any
p. 450 limits on the time one has to listen. Only then, he says, can the music ‘ ow’. Yavit’s statement was

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
typical of the ways rasikas I interviewed spoke about time in North Indian classical music. Like Yavit, other
rasikas often described North Indian classical concerts as having their own, special temporality, distinct
from the temporalities of other aspects of their lives. Their favourite kinds of concerts are those where the
passing of time is not measured: they progress at a slowed-down, leisurely pace, without any hurry or time
limits, possibly lasting many hours. Rasikas report that at such concerts both listeners and also musicians
can lose track of time altogether, leading to experiences of timelessness. Through these conversations, it
became clear that this leisurely way of experiencing time is one of the key sources of value and pleasure that
rasikas derive from North Indian classical music.

The anthropologist Kalpana Ram (2011) has discussed this sense of leisurely time as a feature of classical C21.P13
arts in general (including North and South Indian classical music and also dance) in post-Independence
India. She writes that connoisseurship of all classical arts involves ‘[inhabiting] the time of the present in a
very particular way’: speci cally, she argues, this involves a kind of ‘slowed down’ temporality, in which
aspects of the performance (e.g. kinds of musical improvisation) are geared towards making possible the
experience of ‘savouring’ for connoisseurs in the audience. She links this experience both with ancient
Sanskrit theories of rasa and also with the particular context of post-Independence India (the ‘twenty years
or so’ after Independence in 1947), during which time, she suggests, ‘the leisurely savouring of the arts […]
o ers respite [to the subjects of the new India] from the breathless, compressed temporality of dragging the
nation out of underdevelopment’ (S169). A major contribution of Ram’s article is to draw attention to the
ways in which the arts in India a ord their audiences particular experiences of time (speci cally,
experiences of time as slowed down or leisurely), and she demonstrates the need to situate these
experiences within wider social, political, and historical contexts. However, this article is a largely
theoretical piece; although Ram herself is clearly well-versed in the traditions she discusses, she does not
draw explicitly on any sustained ethnographic or qualitative research with the audiences she theorizes.
Here, building on Ram, I explore the temporality of just one classical tradition (North Indian classical
music) through ethnographic research with a particular group of listeners. Like Ram, I am interested in the
wider social and political consequences of the temporalities that music engenders; however, my
ethnographic work leads me to draw some di erent conclusions from Ram about the ways in which this
temporality operates in the context of contemporary India.

Many aspects of typical North Indian classical performances are well suited to the production of a leisurely C21.P14
sense of time. In part, this emerges from particular qualities of the musical sound (the ‘music itself’).
Classical concerts usually commence with a slow-moving ālāp section, in which the soloist gradually
introduces listeners to elements of the rāg they are performing. This portion of the performance may be
unmetred (in dhrupad) or may occur against the background of an extremely slow metrical cycle (in khyāl);
this section is often described as ‘leisurely’, and the overall attitude to time is relaxed. Rasikas especially
value this part of the performance, in part because of the temporality it engenders. When I asked rasikas
p. 451 about their musical likes and dislikes, many criticized musicians who hurry through these slow, serious
portions of the performance, arriving too soon at the fast, virtuosic portions that normally come towards
the end of a performance; they see this as a sign of a musician who is ‘playing to the gallery’, and as a
symptom of audiences whose attention spans are too short (see also Alaghband-Zadeh 2017a). By contrast,
when musicians dwell on slower sections, rasikas take this as a sign of a high-quality performance. This
corroborates the link that Ram draws between certain kinds of musical improvisation and the experience of
leisurely time in performance (2011: S162). However, the temporality of live performances is not determined
by the music alone; a fuller understanding of musical time needs to take into account how temporalities are
also conditioned by the material a ordances of the performance environment, by the embodied conventions
of performing and listening in this musical culture, and by the broader culture of listening that shapes
rasikas’ engagement with the music.

In addition to qualities of the musical sound, the typical behaviours of both musicians and listeners at live C21.P15

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
performances are also geared towards producing a relaxed approach to time. In India, classical concerts
rarely start on time. Many audience members choose to arrive well after the advertised start time, some
entering after the performance has begun. The music itself often has no clear starting point: most soloists
ease their way slowly into performances. After arriving on stage, they then gradually settle into position,
tuning the instruments in their own time. Even after starting an item, they might stop again to retune or to
ask the sound technician to adjust the mix. If a high-status audience member (such as a senior musician or
well-known patron) arrives, the performers might wait to acknowledge them or to ensure that they are
given a seat towards the front of the audience. As the performance develops, musicians typically pause
between numbers to talk with the audience; during these pauses, audience members start to make requests,
sometimes greatly extending the time of the performance. At longer events and music festivals, audience
members come and go throughout, perhaps taking a break for tea, often served in an adjacent area. These
conventions are in tune with the leisurely atmosphere engendered by some of the most important parts of
the music; they combine to construct performances as spaces that are free from any time pressures.

Nevertheless, not all kinds of performance environment are equally suited to the production of leisurely C21.P16
time. Rasikas are keenly aware of the ways di erent performance spaces engender di erent temporalities.
Many consider the concert hall a less-than-ideal venue for North Indian classical music in part because the
timings there are strictly determined in advance, whereas in private concerts or musical festivals,
performances may proceed without any time limit, perhaps even lasting all night. For example, Yavit
compared the concert hall with the concerts he puts on in his own house:

In concert halls, when you go and listen to music, necessarily so, an artist will be given, allotted C21.P17
some time. Sing for one hour! 45 minutes! One and a half hours! And, unfortunately or fortunately,
people’s attention span in the concert audience is pretty limited, so […] concerts cannot last more
p. 452 than one or two hours. [On the other hand,] I have been involved with my music starting at 9
o’clock at night until 4.30 in the morning!

He notes the requirements such performances place on musicians: they have to have ‘the prowess to sing C21.P18
for hours and hours’. Nevertheless, he said, with the right conditions and the right kinds of listeners, even
musicians come to experience ideal performances as timeless. He told me about how one musician who had
already performed for him reassured a fellow musician who was nervous about the long time slot he gives:
‘You go there, you start singing, the audience and the environment of that evening will see you through for
many hours. [You will not] know where the time has gone.’

Thus, for Yavit, like many other listeners, a key way of distinguishing between good and bad concert C21.P19
environments is to compare the temporalities that emerge from each. A long concert is a sign of both an
excellent performer (someone with ‘prowess’) and also an excellent audience (listeners with the ‘attention
span’ to be able to engage with music over long durations).

Likewise, Sumit (a music organizer and amateur performer) told me about the attitude to time that he C21.P20
cultivates at a music event he organizes. He told me that he customarily gives performers four hours to sing
—far longer than the typical slot in the concert hall. He describes their initial reactions as involving some
trepidation: ‘You know the rst thing that hits the artist is: ‘You’re giving me four hours? I’ve never sung
for this much time. […] How will I manage?’
Nevertheless, he says that they end up managing and even singing for longer than the four hours they are C21.P21
given, ‘because they want to share; they just want to give.’ He links this with the special environment he
tries to create at the event, which he sees as missing in many other musical events, saying, ‘I think it’s just,
like you use the word connect. […] It’s just that feeling of give and take, that openness, which just connects.’
Sumit, like Yavit, drew a link between the ideal environments for performance, a leisurely atmosphere, and
performances that last a long time. For Sumit, this is what facilitates the emotional connection between
performers and audiences.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
Other listeners likewise highlighted the material conditions of di erent performance spaces as more or less C21.P22
conducive to generating an atmosphere of leisurely or timeless time. One rasika told me about the all-night
concerts she used to take her daughter to when she was a child; she told me that her daughter used to listen
until she was tired and then go to sleep on one of the mattresses provided on the oor. As a result, she could
relax and stay all night, making possible some of her most valued musical experiences. Another rasika told
me about his favourite venue for hearing music, saying that he was comfortable there, because there is an
adjacent room serving tea and co ee, meaning that listeners do not have to sit still for hours: ‘As and when
you feel like, you go out and have a cup of co ee, then you come back.’ In each of these cases, an atmosphere
of leisurely temporality is supported in part by the physical environment of the performance space (the
availability of mattresses and co ee). Thus, the slowed-down temporality of North Indian classical
performances cannot simply be reduced to qualities of the music; it emerges from the dynamic relationship
p. 453 between musicians, listeners, the musical sound, performance conventions, and the physical a ordances
of the performance environment.

In an article on experiences of time in a neighbourhood in Lisbon, geographers Paiva et al. (2017) propose C21.P23
that cities o er sets of spatialized ‘temporal resources’, which people use to inhabit time in particular ways.
They show how practices of slowing down or speeding up are tied to (and made possible by) spatial features
of the city. I suggest that their concept of ‘temporal resources’ o ers a useful way of understanding the
temporalities that emerge from musical performances. As my conversations with rasikas revealed, live
musical concerts o er listeners a variety of temporal resources with which to inhabit time. The ‘music
itself’ is one such temporal resource; indeed, rasikas told me that they especially value those elements of the
music that foster a sense of relaxed or leisurely time. However, as Paiva et al. stress, the material
a ordances of particular spaces are also central to the temporalities they make possible. Likewise, live
performances of North Indian classical music o er material and spatial resources which operate in
conjunction with the musical sound in a ording particular experiences of time. Moreover, rasikas’
experiences of leisurely time are made possible by particular conventions of performing and listening, as
well as by the ideologies of musical listening that rasikas invest in and reproduce. Thus, I suggest
understanding North Indian classical concerts as projects of slowed-down temporality, in which sets of
sonic, material, embodied, and discursive resources come together to enable experiences of timelessness
and long duration. As I discuss in the following sections, this slowed-down temporality in turn serves as a
resource for the cultivation of ethical subjects and for listeners to position themselves within the longer
sweep of historical time, occupying a particular position within Indian late modernity.
The Ethics of Being Generous with Time C21.S3

For rasikas, the leisurely temporality of live performances is more than just a source of listening pleasure; C21.P24
this is also infused with ideas about virtue and about what it means to be a good listener. Elsewhere, I have
shown how rasikas’ embodied ways of listening to music constitute a practice of elite connoisseurship: they
are a means through which certain listeners perform a high-status social identity and cultivate themselves

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
as ethical subjects (Alaghband-Zadeh 2017a). Here, I show how this also implicates the ways in which these
listeners inhabit time. I suggest that the leisurely temporality of North Indian classical music is tied to
particular temporal sensibilities, which rasikas adopt in order to foster valued personal characteristics of
generosity and patience and to perform their status as good listeners.

Attitudes to time are a key means through which rasikas distinguish good listeners from bad. For most of the C21.P25
rasikas I have spoken to, patience is a key attribute of a good listener. This, in turn, is an ethical identity,
suggestive of someone with good character. On the other hand, impatience and the inability to sit through
p. 454 slow, serious music are signs that someone is both a bad listener and a bad person. For example, Sumit
told me that a good listener must be ‘a good learner, a patient person. Those are the most important
attributes. Everything else follows.’ Likewise, Arun told me that his involvement in music had led to his
getting to know ‘some very good people’, in part, he said, because listeners of this music have to ‘have
patience’. Statements such as these highlight the ethical signi cance of the slowed-down temporality that
musicians and listeners cultivate at live performances. This temporality makes it possible for individuals to
enact a temporal sensibility characterized by patience, which in turn is a means of performing their status
as both good listeners and also good people.

Beyond the immediate context of the performance, this also applies to much longer durations of time, too. C21.P26
For example, many listeners I interviewed insisted that becoming a rasika or a good listener requires a
commitment of time at a much longer scale: a commitment of many years. As Arun put it, ‘It’s a constant
process. I still feel I haven’t heard enough. […] [Even] if you think that you’ve heard everything, you know
you spend 40 years and, “Shit, look! I’m done!”, it’s never done.’ He continued:

To become a true rasika, you have to have a personal history and it has to be an obsession, or C21.P27
something which goes beyond plain liking of something. […] It’s an attribute which everyone
would like to have, but it don’t come easy. You know human nature is such that we all try to search
for shortcuts and easy way outs and all that, but for anybody passionate about anything there are
no shortcuts.

The language Arun uses here is ethically charged. He constructs the gure of the ‘true rasika’ as a valued and C21.P28
high-status identity (‘an attribute which everyone would like to have’) and a virtuous one, requiring hard
work and the investment of time. This highlights how musical performance and listening make possible the
intersections of temporalities at a variety of di erent scales. For listeners like Sumit and Arun, a willingness
to commit time to music is evidence of virtue, whether it occurs at the scale of the performance or the much
longer scale of a lifetime.

Discourses in which ethical value is place on the act of committing time are not limited to listeners of North C21.P29
Indian classical music; there are similar discourses surrounding musicianship, too. For musicians in this
tradition, time spent practising is an important source of prestige and indicates both personal virtue and a
high status. As with listeners, this occurs at various scales: musicians often tell extraordinary anecdotes
about the number of years they spent training, the number of hours per day their teacher had them play or
sing, the number of months they spent learning one rāg, the number of hours per morning they typically
spend singing one note, and so on. This commitment of time serves a practical purpose: like many musical
traditions, North Indian classical music is di cult and takes many hours of practice and instruction to
master. Dard Neuman (2012) suggests that practising repetitively and at length, even to the point of
boredom, is key to how musicians learn to improvise in this tradition. However, this is also indicative of a
wider ideology, spanning both discourses of listening and musicianship, in which ethical value is placed on
the act of committing time.

p. 455 A leisurely attitude to time characterized my own music lessons with renowned vocalist Sunanda Sharma. I C21.P30
took intensive lessons with Sharma in Delhi in 2008, in preparation for my PhD research. Although I paid

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
Sharma for an hour-long lesson every day, in reality I ended up spending far longer at her house. On a
typical day, I would arrive at the appointed time for one-to-one instruction with Sharma for about 45
minutes. Then other, more advanced students would arrive. I would sit in on their lessons, joining in when
Sharma felt I was able. Sometimes Sharma would choose to see a group of us at once, in order to teach us all
the same composition. After a few hours, Sharma would then send me home for individual practice. There
was never a sense that our lessons were being timed. Rather, the time I spent with Sharma would seem to
move uidly between individual instruction, supervised practice, group instruction with other pupils,
listening in to lessons with other students, running errands, and making tea, sometimes all within a single
visit. Likewise, the workshops Sharma teaches abroad feel similarly open-ended and timeless. During these
workshops, any predetermined timetables are likely to be completely abandoned. We take long pauses from
singing for conversation and Sharma decides when (if at all) we should take breaks for meals or tea based on
how we are doing and the mood in the room. Through these experiences, I gradually learned a new, slowed-
down way of inhabiting time, which I came to associate with North Indian classical music. I found that I
enjoyed the relaxed freedom of learning music without time limits, which contrasted strongly with my prior
experiences learning (and teaching) Western classical music. Without any speci ed end-point looming, it
felt as though there was always space to explore in depth the nuances and subtleties of rāgs and
compositions; I came to view this temporality as a key feature of my education in this music. Moreover, by
inculcating me into a particular attitude towards time, these lessons prepared me to feel at home in the
leisurely environment of live performances and to hold my own among the rasikas around me.

Thus, the temporal sensibility that rasikas cultivate in relation to North Indian classical music is part of a C21.P31
broader constellation of ideas and practices in this tradition; the slowed-down time of the concert is
embedded within a musical culture (and an ethical world) in which individuals, both listeners and
musicians, gain status and prestige by committing time. This listening ideology spans scales, so that being
generous with time is valued both at the scale of the concert and also at the scale of a lifetime. I suggest that
the leisurely temporality that characterizes live performances comes to seem natural in part because it
resonates with these broader sets of values and ideologies that permeate this tradition. Charles Hirschkind
(2006) has famously theorized listening as an ethical practice. In his work on the practice of listening to
cassette-sermons in Egypt, he shows how listeners use these sermons as a way of cultivating ethical
dispositions and ‘an ethically responsive sensorium’ (p. 10). My research with rasikas likewise illustrates
how the act of listening (in this case, to music) can also serve as an act of ethical self-discipline;
importantly, this ethical practice of listening involves ways of inhabiting time. In the moment of
performance, the slowed-down nature of classical concerts makes it possible for rasikas to perform and
p. 456 embody valued personal characteristics of patience and an attitude of generosity with time; this in turn is
a means by which they cultivate themselves as ethical subjects.

Moreover, this attitude to time also has implications in terms of social hierarchies. For the listeners I C21.P32
interviewed, performing the identity of the rasika or ‘good listener’ was not only tied to ideas about personal
virtue; this also identi ed them as part of a respected group of listeners, with a high status within the music
world. More broadly, this also relates to social class. Elsewhere, I have shown how that the construction of
the rasika or good listener of North Indian classical music embodies values of the so-called ‘old middle
class’ in India (Alaghband-Zadeh 2017a). Here, I would like to suggest that this extends to the temporal
sensibilities that rasikas cultivate in performance: rasikas’ understandings of time, like other aspects of their
listening practices, are also congruent with old-middle-class values. The value these listeners place on
being generous with time, or being willing to commit time to music, re ects the value that the old middle
class typically places on deferred grati cation, as opposed to the search for more instant pleasures (see e.g.
Nigam 2004). Moreover, this also illustrates Pierre Bourdieu’s argument that those cultural practices which
require time are best suited to serving as tools of social distinction for elite groups. This is in part because
only members of these groups are likely to have had the free time available to cultivate expertise in these
areas (see e.g. Bourdieu 1984[1979]: 75). In addition, members of elite groups value ‘capacities which,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
requiring a long investment of time […], cannot be acquired in haste or by proxy, and which therefore
appear as the surest indications of the quality of the person’ (281). In the case of India, this approach to time
also indexes the now-disappeared aristocracy, who patronized classical music prior to Independence and
whose behaviour, as Regula Qureshi has argued (2001: 128), served as a model for aspiring middle-class
individuals after Independence. Thus, the temporal sensibilities that rasikas embody are socially productive
at multiple levels, giving rasikas the means to cultivate virtuous forms of the self, to occupy a high-status
position within North Indian classical music, and to reproduce values of an elite group within Indian society
more broadly.

Out of Time: Cultivating Slowness in a Speeded-Up World C21.S4

The immediate temporalities of live performances also relate to the production of time at much longer C21.P33
durations: they are implicated in the way rasikas situate themselves in historical time and in relation to late
modernity in India. Rasikas often draw links between the slowed-down temporality of live performances
and a kind of temporal sensibility which they believe was more prevalent in the past. Most of the older
music-lovers I interviewed spoke with nostalgia about the long or all-night concerts they remembered from
their younger days, lamenting that such concerts are increasingly di cult to come by. Rasikas often put this
p. 457 down to their belief that today’s audiences are increasingly impatient; some highlighted new noise
regulations, according to which all public concerts must nish before 10 p.m. For example, as Arun put it,
‘Those were simple times. People had patience. People were indulgent and […] could sit through three or
four hours.’ For these listeners, the slowed-down temporality of live performances and the patient listening
stance that accompanies it preserve a trace of the past.

Some listeners understood this change in listeners’ attitudes to time in terms of changing attitudes to time C21.P34
more broadly. Ravi reminisced about the all-night concerts he used to go to as a child, explaining this in
terms of a change in attitude that he has observed in both musicians and listeners:

Concerts typically started at 8 o’clock in Delhi, and went on as long as they went on, because the C21.P35
audiences sat as long as musicians were willing to sing and musicians sang as long as the
audiences were willing to listen.

Ravi understood this changing attitude to concert length in terms of a broader shift in temporal C21.P36
sensibilities. He described how this relaxed attitude to time also applied to his music lessons:

In those days, it wasn’t a by-the-hour kind of thing. The whole relationship was very informal. On C21.P37
days I would tell my teacher, ‘Achha ji [OK, sir], today I don’t feel like learning, so we’ll play cricket
today. So Master-ji [the teacher] would play cricket with me in the garden for two hours, for an
hour, hour and a half, come back, have tea and then he would say, ‘OK, I’ll play cricket on the
condition that you play for half an hour.’ So you learn for half an hour. So that’s how it went.

In this description, Ravi evokes a past world where time was not measured ‘by the hour’. Note the C21.P38
aristocratic context also implied by this description: in telling me this story, Ravi here makes it clear that he
grew up in a family context where they could a ord to keep a music teacher employed as part of the
household sta , constantly on hand to teach the children of the house. Thus, rasikas like Ravi use the
slowed-down temporality of contemporary performances in order to imagine and experience a way of
inhabiting time which they understand as indicative of the past. In such descriptions, rasikas drew links
between temporalities at very di erent scales: the immediate temporality of live performances is implicated
in how they understand the sweep of time at a much larger scale, including changes that they describe as
taking place over a number of decades.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
These understandings of musical time as associated with the past form part of a broader set of discursive C21.P39
tropes through which rasikas situate the leisurely time of North Indian classical music as set apart from the
temporal logics of modernity. For example, Meghna drew a contrast between the temporality of North
Indian classical music and the everyday realities of contemporary life in India. She compared the music
world with what she called the ‘the world we live in’, which she described as materialistic and goal-
oriented, saying:

p. 458 I think we [in India] are in some weird state of trying to become part of a global world. C21.P40
Everything’s become more generic and also there’s this insane drive towards material
achievements, whereas people don’t realize that you have to balance your life out really. […] The
world we live in [is] very limited nowadays. We’re driven by an external view of success and
material goods and things that we think will make us happy. And it’s in a sense also a Western
model of what is good for you.

On the other hand, she described music as ‘a whole second parallel universe’: C21.P41

It was a very amorphous world that had no rational structure, the way my formal education or the C21.P42
way my other life, my job etc. had. So this was a world that was very timeless and there were no
goals you could really achieve as such; it was just something that you dipped into an ocean and it
was an in nite space to explore. […] It’s just been a great anchor for me, because it’s a very
beautiful space that remains unchanged through my life.

Note how Meghna constructs the music world here as a realm removed from modernity, but also a realm C21.P43
with its own temporality. Whereas the ‘world we live in’ is materialistic and goal-oriented, characterized by
an ‘insane drive towards material achievements’, the music world, for her, is ‘timeless’, unchanging, and a
space to explore, rather than a place of goals. Thus, she nds in music a kind of temporal sensibility which is
opposed to, and also an antidote to, the goal-oriented temporality of contemporary India.

Likewise, Sumit also construed music as a realm set apart from the contemporary world, requiring a C21.P44
di erent temporal sensibility in order to appreciate it properly. He lamented that modern life is not
conducive to the best kinds of musical experience, saying that it is in danger of making people’s experiences
‘di erent’ (i.e. less good). In doing so, he evoked a common discursive trope that links modern technology
with distraction and being in a rush:

Everything depends on the mood you are in, with the receptivity of [the audience]. If you are in a C21.P45
rush to get back home or to catch the last train back, it’s a little di erent. [If] there’s too much
activity happening or you have the modern gadgets to disturb you and distract you, and want to get
distracted, it’s a di erent experience.

For him, the solution was to organize residential musical events in a location that he describes as set apart C21.P46
from modernity, including, crucially, from the normally ever-present distraction of mobile telephones:
So I have these artists who are performing for me and I host it in a place which is really cut away C21.P47
from Mumbai. It’s desolate, completely insulated from the city, from civilization so to say, amidst
mountains. So we just park ourselves for three days. Mobile phones don’t work. And it’s only for
the chosen few who want to cut away from everyday life and want a slice of this beautiful art.

p. 459 This description is indicative of how musical time is shaped by elements that go well beyond the musical C21.P48
sound: here, the temporalities of musical experience are shaped by particular spaces and by the presence or

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
absence of particular kinds of technology. Note also the way this understanding of time is intertwined with
processes of social distinction, as discussed in the previous section. (This suggests that the construction of
rasikas as virtuous subjects is in part related to practices of occupying a particular position in relation to
modernity; I shall return to this idea later.)

Throughout my interviews with music-lovers, other listeners too construed music as at odds with the C21.P49
temporality of modern life. In their descriptions, India today is a place of noise regulations, tra c, long
work hours, short attention spans, ever-present and potentially distracting technology (especially mobile
phones). They emphasized that all of these get in the way of good music-making. So the leisurely
temporality of North Indian classical concerts is tied both to ideas about historical time and also to a sense
of alienation from contemporary life: the slowed-down temporality of rasikas’ listening experiences serves
to situate listeners within the time of the present, occupying a position of alienation in relation to
modernity.

Rasikas’ experiences therefore o er insights that can contribute to current scholarly discussions about time C21.P50
and modernity. Many of these discussions have centred on ideas of acceleration: they have probed the sense
that modernity is speeded up, often linking this with the emergence of new technologies (e.g. Rosa 2013;
Rosa and Scheuerman 2010; Scheuerman 2004; Tomlinson 2007). This sense that the present is
characterized by speed while the past represents relative slowness might feel intuitively real; however,
various scholars have cautioned against uncritically assuming that speed is the only (or even the main)
temporal logic at play in contemporary life. Keightley invites a move ‘beyond a one-dimensional
characterization in which speed and immediacy monopolize accounts of how time is encountered and lived’
(2012: 4); as she writes elsewhere, ‘Assessments of mediated time as characterized by speed and
acceleration ignore the variety of temporal experiences that are a orded by media technologies’ (2013: 71).
This work forms part of a broader body of literature that has examined the multiple temporalities that
coexist in contemporary contexts. This has included research on contemporary practices through which
people cultivate stillness or slowness (such as the slow food movement), often in opposition to ideas of
acceleration and speed (e.g. Bissell and Fuller 2011). These projects of stillness confound the idea that the
present is uniformly characterized by acceleration. Rasikas’ listening practices are one such project.
Through their listening practices (and the discourse and ideology surrounding them), rasikas construct
North Indian classical music as a space of slowed-down time, in contrast to what they understand as a
speeded-up modern world. In this context, discourses of acceleration have a powerful social reality, even
though rasikas’ own practices complicate the idea that modernity is characterized primarily by speed.
Moreover, this case study reveals how ideas about both acceleration and also slowness can be socially
productive: rasikas draw on these ideas in order to craft a social identity for themselves as set apart from
modernity.

p. 460 As well as forming part of a global set of practices of cultivating slowness, the temporality that rasikas C21.P51
inhabit at live performances also performs more speci c cultural work in the context of contemporary India.
Anthropologist Purnima Mankekar has shown how public discourse in India in the early twenty- rst
century has constructed the nation as a place of aspiration and forward drive. She highlights a set of
discourses which she labels ‘Aspirational India’: ‘Aspirational India engages a temporality of self-
advancement and futurity; it is about overcoming a sense of inferiority based on a sense of a temporal lag
vis-à-vis the so-called developed world’ (2015: 227). By contrast, rasikas’ listening practices resist the sense
of future orientation and aspiration that characterize contemporary discourse on the nation; instead, they
use music to position themselves in a space that is located outside of modernity. In their discussions of
music and listening, rasikas articulated a sense of alienation in relation to contemporary life; they were far
more likely to talk optimistically about the past than the future. Rasikas’ listening practices are thus a way
for them to set themselves apart from the certain dominant discourses on the nation, and to refuse to take
up the identities these discourses a ord.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
Rasikas’ backward-looking stance resonates with a wider middle-class moral panic about modernity and C21.P52
consumer culture, which is played out in part through attitudes to time. This moral panic emerges out of
broad economic and cultural changes in India over the last few decades, including economic liberalization
from the 1980s onwards (and especially in the 1990s), making available a range of new commodities and
coinciding with a large-scale move away from post-Independence, Gandhian principles of austerity and
asceticism. In her work on attitudes to consumption in an Indian city (Baroda), Margit van Wessel (2004)
shows how middle-class people articulate a sense of ambivalence regarding consumer culture. Although
they accept that the consumption of new commodities is now an important part of their lives, they worry
about its morality, seeing it as a form of ‘debased materialism’, in which ‘people seek self-realization or
self-expression through goods rather than through spiritual or social pursuits, which leads to the
evaluation of individuals on the basis of their material possessions rather than other (higher) aspects of
their person’ (p. 95). Moreover, these attitudes to consumption have a temporal aspect. Van Wessel’s
participants constructed the past as ‘simple’ (pp. 96, 99); they attached virtue to people who ‘organize their
lives in a sober, digni ed manner, live frugally and seek neither enjoyment nor status through
consumption’ and they equate being ‘good’ with being ‘non-modern’ (p. 99). She notes that her
participants thus echo the ‘view, consistently found in elite scholarly and journalistic circles in India, that
the middle class is in the grip of a moral crisis due to the advent of consumer culture’ (p. 108). For rasikas,
crafting their identity as listeners in opposition to modernity is, likewise, a way of eschewing the identity of
the reckless consumer (a dominant trope in contemporary public discourse in India), instead occupying a
more virtuous identity, associated with the past. This adds another layer of signi cance to the construction
of the ‘good listener’ or rasikas as a virtuous identity, as discussed in the previous section, suggesting that
the sense of virtue attached to listening is also in part a matter of how rasikas position themselves in time
and within these wider contemporary discourses.

p. 461 This case study therefore illustrates the importance of understanding projects of slowness within the C21.P53
speci c contexts in which they occur. Rasikas’ practices of cultivating slowness through music are a means
by which they construct and situate themselves within imagined histories of Indian classical music and,
more broadly, histories of the nation. They are a means for rasikas to craft and occupy positions of virtue
within contemporary Indian society. This project of slowness resonates with numerous similar projects that
have appeared in various global contexts; however, it also performs important social and cultural work in
situ, relating to local discourses about modernity, virtue, and the nation.
Conclusion: Music as a Temporal Resource C21.S5

Sociologist Tia DeNora has theorized music as a ‘resource for producing social life’ (2000: 129). My C21.P54
ethnographic work with expert listeners of North Indian classical music (so-called rasikas) suggests that
this extends to the social construction of time: that is, that music is an example of what Paiva et al. call a
‘temporal resource’ (2017). In this chapter, I have shown how music is a powerful resource for rasikas to

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
inhabit time at various scales. In the moment of performance, listening to music is a way for these listeners
to cultivate experiences of timelessness or of leisurely time. They situate these experiences in contrast to
the temporalities of everyday life: while they understand time outside music contexts as measured and
speeded up, rasikas use music to experience the passage of time as unmeasured and slowed down. This
temporality is an important source of listening pleasure, but it also has social consequences: it is a means by
which rasikas inhabit the identity of the ‘good listener’, an identity which is infused with virtue and which is
associated with a high social status both within and beyond music contexts. Listening to music also serves
as a temporal resource for rasikas to inhabit time at much longer scales: rasikas draw on their experiences of
leisurely time in music in order to situate themselves within imagined histories and to occupy a position
that they understand as set apart from the temporal logics of late modernity in India.

Theorizing music as a temporal resource has signi cant advantages for the study of music and time. This C21.P55
avoids the pitfalls of collapsing musical temporalities into the qualities of the musical sound or score,
instead making possible multidimensional analyses of the ways in which musical practices are implicated in
the social construction of time. For example, as I have highlighted here, this approach makes it possible to
look beyond the musical sound, showing how music operates in conjunction with the variety of spatial,
material, and technological resources that exists at any site of performance or listening; listeners use these
resources together in order to make possible particular experiences of time. Still, this approach does not
p. 462 preclude musical analysis. On the contrary, music analysis remains an invaluable tool for understanding
the precise nature of the sonic a ordances that musical sound o ers its listeners.

As DeNora stresses, understanding music as a resource acknowledges the agency of listeners to use music in C21.P56
a variety of creative and individual ways. DeNora is less interested in how people’s engagement with music
might also be determined by ideology, cultures of listening, or the large-scale structures of power in which
listeners are embedded; however, her theoretical model is well-suited to analysis along these lines. Here,
for example, I have highlighted ways in which rasikas’ listening practices are conditioned by social
hierarchies, the expectations of a particular culture of listening, and by the broader discursive and
ideological contexts in which their listening occurs. This approach also makes it possible to draw links
between musical temporalities and wider temporalities of nation and modernity without simply relying on
similarities or homologies (as in structuralist approaches to musical time). Instead, I have shown how
listeners draw on their musical experiences in order to construct larger narratives about changing times and
temporalities. While this may at times be grounded in relationships of similarity, this can also involve other
types of relationships, including contrast or di erence. For example, rasikas’ experiences of slowness in
music contribute to their understandings of modernity as speeded up.

Rasikas’ listening practices illuminate the multiplicity of intersecting temporalities that emerge out of sites C21.P57
of musical performance and listening. The immediate, moment-to-moment temporalities of live
performance contribute to the production of temporalities at much longer scales; listeners’ experiences of
time in the moment of performance are enmeshed in broader temporalities of modernity, social class, the
nation, and imagined histories. Moreover, I suggest that the experience of music in performance is part of
what makes these larger-scale temporalities a ectively real. Listeners draw on their immediate, a ective,
embodied experiences of time in performance in order to situate themselves in relation to imagined
histories and to craft and occupy particular positions within Indian late modernity. Listening is thus a
complex temporal practice, implicating and intertwining temporalities at various scales, as well as
discourses, dispositions, modes of embodiment, and social formations. Above all, I have demonstrated the
importance of ethnography (and ethnography with listeners in particular) for understanding the complex
sets of relationships between music and time. Ethnography is a crucial tool for uncovering the sociality of
musical temporalities: how time is made by people, and how it is embedded within the a ective fabric of
musical experience, and yet also fully part of the social world.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
Notes
1. Lorraine Plourdeʼs work on ways of listening to the Japanese experimental genre onkyō is an exception to this; I draw C21.N1
heavily on her approach in this chapter. There, based on ethnographic work with listeners, she writes about how the
experience of listening to onkyō involves ʻa heightened awareness of space and the sensation of timeʼ (2008: 286).
p. 463 However, time is not a major focus of her argument. Ethnography has also formed part of the diverse methodological
toolkit employed in important new work on entrainment and musical listening, especially work by Martin Clayton, Laura
Leante, and Glaura Lucas (see e.g. Clayton 2008; 2013; Lucas et al. 2011; Lucas 2013). However, this work tends to focus on
the cognitive aspects of how listeners experience time in the moment of performance, whereas I am more interested in
the cultural construction of time at multiple scales and how this emerges from the intertwining of discourse and
embodied practice.

2. I quote a number of these interviews here; in all cases, I have anonymized these sources fully, by replacing intervieweesʼ C21.N2
names with pseudonyms and not dating/placing the interviews.

3. This chapter forms part of a wider project on this musical scene (see also Alaghband-Zadeh 2017a; 2017b). C21.N3

4. As Amita Baviskar and Raka Ray have pointed out, despite the implication that middle-class people represent ʻthe C21.N4
everymanʼ, middle-class groups in India in fact hail from towards the top of the income spectrum (2011: 2, 7–8).

5. For more on this culture of listening, see Alaghband-Zadeh (2017a; 2017b), Clayton (2007), Clayton and Leante (2015), C21.N5
Leante (2016), Neuman (1990[1980]), and Silver (1984).
References C21.S6

Alaghband-Zadeh, C. (2017a). Listening to North Indian classical music: How embodied ways of listening perform imagined C21.P58
histories and social class. Ethnomusicology 61(2): 207–233. https://doi.org/10.5406/ethnomusicology.61.2.0207
Google Scholar WorldCat

Alaghband-Zadeh, C. (2017b). Still, silent listening in India: The meanings of embodied listening practices. In H. Barlow and C21.P59

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
D. Rowland (eds), Listening to music: People, practices and experiences. http://ledbooks.org/proceedings2017/#sec_189_h1
Google Scholar Google Preview WorldCat COPAC

Bar-Yosef, A. (2001). Musical time organization and space concept: A model of cross-cultural analogy. Ethnomusicology 45(3): C21.P60
423–442. https://doi.org/10.2307/852865
Google Scholar WorldCat

Baviskar, A., and Ray, R. (eds). (2011). Elite and everyman: The cultural politics of the Indian middle classes. Routledge. C21.P61
Google Scholar Google Preview WorldCat COPAC

Becker, J. (1979). Time and tune in Java. In A. L. Becker and A. A. Yengoyan (eds), The imagination of reality: Essays in Southeast C21.P62
Asian coherence systems, 197–210. Ablex.
Google Scholar Google Preview WorldCat COPAC

Berger, K. (2008). Bachʼs cycle, Mozartʼs arrow: An essay on the origins of musical modernity. University of California Press. C21.P63
Google Scholar Google Preview WorldCat COPAC

Bergson, H. (1911). Creative evolution. Henry Holt. C21.P64


Google Scholar Google Preview WorldCat COPAC

Bissell, D., and Fuller, G. (eds) (2011). Stillness in a mobile world. Routledge. C21.P65
Google Scholar Google Preview WorldCat COPAC

Born, G. (2015). Making time: Temporality, history, and the cultural object. New Literary History 46(3): 361–386. C21.P66
https://doi.org/10.1353/nlh.2015.0025
Google Scholar WorldCat

Bourdieu, P. (1984). Distinction: A social critique of the judgement of taste. Routledge & Kegan Paul. C21.P67
Google Scholar Google Preview WorldCat COPAC

Clayton, M. (2000). Time in Indian music: Rhythm, metre, and form in North Indian rāg performance. Oxford University Press. C21.P68
Google Scholar Google Preview WorldCat COPAC

Clayton, M. (2007). Time, gesture and attention in a khyāl performance. Asian Music 38(2): 71–96. C21.P69
https://doi.org/10.1353/amu.2007.0032
Google Scholar WorldCat

Clayton, M. (2008). Toward an ethnomusicology of sound experience. In H. Stobart (ed.), The new (ethno)musicologies, 135–169. C21.P70
Scarecrow Press.
Google Scholar Google Preview WorldCat COPAC

p. 464 Clayton, M. (2013). Entrainment, ethnography and musical interaction. In M. Clayton, B. Dueck, and L. Leante (eds), Experience C21.P71
and meaning in music performance, 17–39. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Clayton, M., and Leante, L. (2015). Role, status and hierarchy in the performance of North Indian classical music. C21.P72
Ethnomusicology Forum 24(3): 414–442. https://doi.org/10.1080/17411912.2015.1091272
Google Scholar WorldCat

DeNora, T. (2000). Music in everyday life. Cambridge University Press. C21.P73


Google Scholar Google Preview WorldCat COPAC

Hirschkind, C. (2006). The ethical soundscape: Cassette sermons and Islamic counterpublics. Columbia University Press. C21.P74
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
Hyland, A. M. (2016). In search of liberated time, or Schubertʼs Quartet in G Major, D. 887: Once more between sonata and C21.P75
variation. Music Theory Spectrum 38(1): 85–108. https://doi.org/10.1093/mts/mtv023
Google Scholar WorldCat

Johnson, J. (2015). Out of time: Music and the making of modernity. Oxford University Press. C21.P76
Google Scholar Google Preview WorldCat COPAC

Kapchan, D. (2015). Body. In D. Novak and M. Sakakeeny (eds), Keywords in sound, 33–44. Duke University Press. C21.P77
Google Scholar Google Preview WorldCat COPAC

Keightley, E. (2012). Time, media and modernity. Palgrave Macmillan. C21.P78


Google Scholar Google Preview WorldCat COPAC

Keightley, E. (2013). From immediacy to intermediacy: The mediation of lived time. Time and Society 22(1): 55–75. C21.P79
https://doi.org/10.1177/0961463X11402045
Google Scholar WorldCat

Leante, L. (2016). Observing musicians/audience interaction in North Indian classical music performance. In I. Tsioulakis and C21.P80
E. Hytonen (eds), Musicians and their audiences: Performance, speech and mediation, 34–49. Routledge.
Google Scholar Google Preview WorldCat COPAC

Lucas, G. (2013). Performing the rosary: Meanings of time in Afro-Brazilian Congado music. In M. Clayton, B. Dueck, and L. Leante C21.P81
(eds), Experience and meaning in music performance, 86–107. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Lucas, G., Clayton, M., and Leante, L. (2011). Inter-group entrainment in Afro-Brazilian Congado ritual. Empirical Musicology C21.P82
Review 6(2): 75–102. https://doi.org/10.18061/1811/51203
Google Scholar WorldCat

Mankekar, P. (2015). Unsettling India: A ect, temporality, transnationality. Duke University Press. C21.P83
Google Scholar Google Preview WorldCat COPAC

McClary, S. (2000). Temp work: Music and the cultural shaping of time. Musicology Australia 23(1): 160–175. C21.P84
https://doi.org/10.1080/08145857.2000.10415918
Google Scholar WorldCat

McGraw, A. C. (2008). Di erent temporalities: The time of Balinese gamelan. Yearbook for Traditional Music 40: 136–162. C21.P85
Google Scholar WorldCat

Murphy, S. (ed.). (2018). Brahms and the shaping of time. University of Rochester Press. C21.P86
Google Scholar Google Preview WorldCat COPAC

Neuman, D. (2012). Pedagogy, practice, and embodied creativity in Hindustani music. Ethnomusicology 56(3): 426–449. C21.P87
https://doi.org/10.5406/ethnomusicology.56.3.0426
Google Scholar WorldCat

Neuman, D. M. (1990). The life of music in North India: The organization of an artistic tradition. Manohar. C21.P88
Google Scholar Google Preview WorldCat COPAC

Nigam, A. (2004). Imagining the global nation: Time and hegemony. Economic and Political Weekly 39(1): 72–79. C21.P89
Google Scholar WorldCat

Paiva, D., Cachinho, H., and Barata-Salgueiro, T. (2017). The pace of life and temporal resources in a neighborhood of an edge C21.P90
city. Time and Society 26(1): 28–51. https://doi.org/10.1177/0961463X15596704
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471006 by National Science & Technology Library user on 26 May 2023
Plourde, L. (2008). Disciplined listening in Tokyo: Onkyō and non-intentional sounds. Ethnomusicology 52(2): 270–295. C21.P91
Google Scholar WorldCat

Qureshi, R. B. (2001). In search of Begum Akhtar: Patriarchy, poetry, and twentieth-century Indian music. World of Music 43(1): C21.P92
97–137.
Google Scholar WorldCat

p. 465 Ram, K. (2011). Being ʻrasikasʼ: The a ective pleasures of music and dance spectatorship and nationhood in Indian middle-class C21.P93
modernity. Journal of the Royal Anthropological Institute 17: S159–S175. https://doi.org/10.2307/23011430
Google Scholar WorldCat

Rosa, H. (2013). Social acceleration: A new theory of modernity. Columbia University Press. C21.P94
Google Scholar Google Preview WorldCat COPAC

Rosa, H., and Scheuerman, W. E. (eds) (2010). High-speed society: Social acceleration, power, and modernity. Penn State C21.P95
University Press.
Google Scholar Google Preview WorldCat COPAC

Rowell, L. (1979). The subconscious language of musical time. Music Theory Spectrum 1: 96–106. https://doi.org/10.2307/745781 C21.P96
Google Scholar WorldCat

Scheuerman, W. E. (2004). Liberal democracy and the social acceleration of time. Johns Hopkins University Press. C21.P97
Google Scholar Google Preview WorldCat COPAC

Silver, B. (1984). On the adab of musicians. In B. D. Metcalf (ed.), Moral conduct and authority: The place of adab in South Asian C21.P98
Islam, 315–329. University of California Press.
Google Scholar Google Preview WorldCat COPAC

Taylor, B. (2016). The melody of time: Music and temporality in the Romantic era. Oxford University Press. C21.P99
Google Scholar Google Preview WorldCat COPAC

Tomlinson, J. (2007). The culture of speed: The coming of immediacy. Sage. C21.P100
Google Scholar Google Preview WorldCat COPAC

Wessel, M. van. (2004). Talking about consumption: How an Indian middle class dissociates from middle-class life. Cultural C21.P101
p. 466 Dynamics 16(1): 93–116. https://doi.org/10.1177/0921374004042752
Google Scholar WorldCat
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
CHAPTER

22 C22Timing in Palaran: Coordination, Control, and


Excitement in Javanese Collaborative Vocal Accompaniment

Jonathan Roberts

https://doi.org/10.1093/oxfordhb/9780190947279.013.25 Pages 467–484


Published: 08 December 2021

Abstract
Palaran are elements within gamelan repertoire that are derived from the melodies used to recite texts
written in Javanese poetic metres. When used as palaran within gamelan performance (rather than in
their original form as poetic recitation, or macapat), these melodies are supported and constrained by a
metrically xed structure of core instrumental notes and surrounded by a web of spontaneous melodic
accompaniment, involving multiple musicians. This chapter explores the complex ways in which
musicians control, negotiate, and coordinate timing in this exible yet precise form of musical
interaction. It examines the rules of ideal cohesive performance, the transmission of strategies for
successfully learning how to achieve this, and what can be learnt from occasions when palaran go
wrong and the coordination of timing goes awry. It then argues that the sense of risk involved in
managing timing in these ways is a signi cant part of what makes palaran one of the most popular
elements within the gamelan repertoire.

Keywords: Javanese gamelan, coordination, recitation, negotiation, paralan pedagogy


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Introduction C22.S1

THIS chapter explores the issue of coordinating timing in the context of Central Javanese Classical gamelan. C22.P1
It does this by examining how the timing rules for unaccompanied, melodic recitation of poetry are
reworked when the same melodies are used in conjunction with instrumental accompaniment as a showcase
for vocal skill and ensemble coordination. One of the most popular elements within the Javanese classical

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
gamelan repertoire (karawitan) is the palaran. A single vocalist is given the space to display their skill by
singing highly ornamented and rhythmically exible melodic lines, supported by a pared-down
instrumental accompaniment which weaves a supportive foundation of pulse and pitch. The level of
popularity of palaran can be seen in the number of recordings which speci cally focus on this form of
performance. In a non-theatrical performance setting, lasting anywhere between three and eight hours,
three or four palaran, lasting about 20 minutes in total, might occur. There are, however, a large number of
entire 70-minute albums on cassette and compact disc dedicated to palaran, which contain no other musical
material apart from brief moments of srepegan, the generic drama accompaniment genre which frames
palaran. This is the karawitan equivalent of entire jazz albums consisting only of solos without the original
piece they are working within, or collections of cadenzas without the concerto in which they originally
occur.

p. 468 I argue that this level of popularity derives, at least in part, from the excitement generated by a level of C22.P2
tension in the coordination of timing. This tension can be perceived by audiences for two reasons. First, the
basic melodic contours and texts of palaran are well known as a result of their origins in standard poetic
recitation (macapat), and thus the ways in which they are exed in performance are readily felt by a majority
of listeners. Secondly, this sense of risk in coordination is not otherwise a major feature of karawitan;
although the female singer (pesindhèn) usually arrives at structural pitches after the rest of the group has
played them, there is no sense of risky coordination in this, simply the pleasure of a nely crafted delay
against the regular progression of the instrumental ensemble. In almost all other cases the normative rules
of playing karawitan emphasize well-controlled, smooth, and practically imperceptible coordination.

The distinctive nature of palaran is also evident in the fact that in the syllabus of the Indonesia Performing C22.P3
Arts Academy (ISI) in Surakarta (from my personal experience in 2000 and 2012) an entire semester of the
vocal curriculum focused on palaran and the srepegan in which they are framed, with even the most
resolutely unenthusiastic singers expected to memorize and perform at least four of these mini-suites.
Palaran are also played almost exclusively by professional musicians or by amateur groups which are led by
a professional and contain large numbers of semi-professional players. During my research on non-
commercial gamelan associations in Surakarta (Roberts 2015) I played with fourteen such groups, and only
in one were palaran even attempted. The others did not even play the repertoire for accompanying shadow
puppetry, such as srepegan, which acts as the springboard from which palaran are launched.

In this chapter, I rst introduce macapat, which is the source for the material used in palaran, setting out the C22.P4
ways in which timing is regulated in this unaccompanied form of melodic recitation. I then explore the ways
in which timing has to be approached di erently when this material is performed with multiple players
accompanying it, and the speci c moments when coordination of timing becomes crucial. An important
factor in this exploration is looking at examples of what can cause a less than optimal performance of a
palaran to go wrong, taking seriously the importance of mistakes and failure in understanding musical
experience. The ways in which palaran can break down throw light on what needs to be in place for a
successful rendition. I also include discussion of how palaran performance is taught. The issues of timing
discussed here mean that it is one of the few elements within karawitan where, in my experience, teachers
explicitly acknowledge this di culty and give hints and instructions as to how to negotiate potential
pitfalls. It is also one of the few elements of karawitan where musicians will explicitly and openly assist or
correct other players in rehearsal and occasionally even performance.
The material on which this chapter is based is primarily ethnographic, being drawn from my experience C22.P5
learning and performing gamelan, and speci cally macapat and palaran, in Surakarta, Central Java. This
includes two years between 1999 and 2001 spent as a student at ISI in Surakarta (informally referred to as
Solo), where I rst learnt to sing and accompany palaran and took part in a local informal macapat session,
p. 469 and a year of eldwork in 2012 spent rehearsing and performing with 14 amateur groups in the same city,
when I focused more heavily on macapat and became part of the team responsible for reciting in front of the
Sultan’s residence on Thursday nights. As a result, the material is based entirely on Solonese practice: I have

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
no direct experience of macapat in Yogyakarta, or of rambangan, which is the genre roughly analogous to
palaran in that city. My learning and playing experience includes formal academy classes, individual tuition
on various instruments and voice, and extensive rehearsal and performance with a large number of groups
in Java. The description here focuses almost entirely on timing; Brinner (1995: 234–244) provides another
detailed description of palaran, but his description and analysis address broader aspects of interaction in the
context of examining how musical knowledge is acquired, demonstrated, and negotiated by Javanese
musicians.

This chapter aims to provide an example of how implicit rules of timing are put under strain and C22.P6
renegotiated under the constraints of interactive performance. The di culties encountered when moving
from performing a speci c text and melody as macapat to performing it as a palaran show how multiple
approaches to structuring time for the generation of tension and resolution are brought together in a
context where individual musicians are required to take more immediate decisions than usual whilst also
being required to t these decisions to those being taken by others. In the terms deployed by Do man (2011:
204), the relationship between tacit musical knowledge and musical conduct is put under greater strain in
the performance of palaran than is usually the case in karawitan. More focus is placed on issues of timing
because the structure of palaran generates a situation where the musical knowledge implicit in each
musician’s role cannot be applied automatically but must be worked out moment to moment in response to
others.

An Explanation of the Notation System Used Here C22.S2

The examples of gamelan music that I provide in this chapter are presented using the kepatihan form of C22.P7
notation, the system now used throughout Java. It is essentially a form of cypher notation which shares
many features with the Galin–Paris–Chevé system and the cypher notation system based on it, which is
now used across East Asia. It uses numbers to represent the pitches of the scales. In order to clarify further
for voices, or instruments with a range of more than one octave, dots are placed below or above the number
to represent lower or higher pitches. This system of dots always represents the relative pitches of one voice
or instrument, rather than overall relative pitch. Thus, both men and women sing from exactly the same
notation but automatically render the pitch at the relative position within their own range. There are many
speci c features of kepatihan which only occur in instrumental notation, and so here I only deal with those
elements which are relevant to vocal notation. The metrically consistent ow of notes in instrumental
Javanese gamelan does not require a notation system that provides detailed information about rhythmic
p. 470 timing. In metrically consistent vocal notation, when more than one note occurs in the space of one beat,
a bar over the two notes is used to mark this, with further subdivisions attracting additional bars. All of the
groups of notes shown here, for example, occur within the space of one beat:

C22.P8

Accurately representing non-metric movement in vocal parts is, however, somewhat more di cult. C22.P9
Generally speaking, the ow of rhythmic movement is shown by spacing on the page, with a larger gap
between notes indicating a longer time interval between them and vice versa. The barring system
mentioned above is also sometimes used to indicate particular rhythmic closeness between two pitches.
Notes that are sung to the same syllable are grouped with underlining. In addition to the basic number
notation, other symbols are used to show the places where the structural instruments play, and these will be
particularly important in discussions of palaran. Strokes on the largest hanging gongs (gong ageng and gong
suwukan), the horizontally mounted pot gongs (kenong), and the smaller mid-range gongs (kempul) are
marked in the following way if they fall on a pitch 6:

C22.P10

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
Macapat C22.S3

Formal written material in Javanese before Indonesian independence in 1945 was composed almost C22.P11
exclusively in poetic metre, including personal letters, court bulletins, and speci cally poetic literature
(Arps 1992: 3). This use of metre brought with it an intimate connection to music: traditionally this material
was not read internally and silently but was read aloud, and this reading made use of a repertoire of
recitation melodies which were associated with individual metres and their variants (Arps 1992: 5). This
means that a written text in Javanese is pregnant with the potential for musical expression.

The features which de ne the di erent metres are the number of lines in a stanza, the number of syllables C22.P12
in each of those lines, and the nal vowel of each line. The number of syllables in a line may be the same
throughout or may vary signi cantly. The kinanthi metre, for example, consists of six lines, all of eight
syllables, with the nal vowels being u, i, a, i, a, and i (Darnawi 1982):

Na-li-ka-ni-ra ing da-lu C22.P13


Wong a-gung mang-sah se-mè-di C22.P14
Si-rep kang ba-la wa-na-ra C22.P15
Sa-da-ya wus sa-mi gu ling C22.P16
Na-dyan a-ri su-dar-sa-na C22.P17
Wus da-ngu nggè-ni-ra gu-ling C22.P18

(Saprodjo 2002) C22.P19

p. 471 The pucung metre, which will be discussed further, consists of four lines of twelve, six, eight, and twelve C22.P20
syllables respectively, ending with u, a, i, and a (Darnawi 1982):

Ngèl-mu i-ku ke-la-ko-né kan-thi la-ku C22.P21


Le-ka-sé la-wan khas C22.P22
Te-ge-sé khas nyan-to-sa-ni C22.P23
Se-tya bu-dya pa-nge-ke-sé dur ang-ka-ra C22.P24

(Saprodjo 2002) C22.P25

The communal recitation of this canon of literature, particularly those books attributed to members of the C22.P26
royal families of Surakarta and Yogyakarta and the court poets, is a key component of what contemporary
Javanese people think of as constituting traditional culture, and also forms an important part of training for
musicians. In Surakarta and Yogyakarta there are still regular meetings of dozens of people coming together
on a monthly basis to work their way through the most respected texts. The memorization of certain
exemplary macapat and recitation melodies remains a part of the primary school syllabus in Central Java
under the title of ‘local vocal art’ (seni suara daerah), and there are still regular competitions for recitation
across a wide gamut of age ranges at local and provincial level. There are also speci c melodies for reciting
some key texts which are only allowed to be rehearsed or used in recitation within the walls of the palace
compounds. The words of the text are regularly used outside the palaces, but with other melodic contours
which are not restricted in this way. Despite the intimate links between melody and text, gathering to recite
macapat is still considered to be a form of reading, in contrast to the use of the same texts and melodies as
palaran, where they are explicitly classi ed as musical.

Timing in Macapat Recitation C22.S4

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
The recitation of macapat follows conventional patterns of timing which allow for individual variation and C22.P27
interpretation while maintaining a basic level of consistency: these patterns relate to the ow of syllables
within a line and the use of pacing to draw attention to important pitches. One of the most fundamental
rules is encoded in the name macapat, which is constructed from the two sense units maca, meaning to read,
and pat, which is an abbreviated form of papat, meaning four. During recital, groups of four syllables are
phrased as discrete units, with distinct pauses between them. This structuring of time overrides the breaks
between words, and it is quite possible for there to be pauses between the syllables of a word in one verse
but for the pause to fall between words in another. In some metres, where the syllable length of lines is not
divisible by four, there are still conventional places within the line where these pauses occur. The rhythmic
and structural emphasis of Javanese melodies is always moving towards the nal note. This is referred to as
the sèlèh note, the note where the melody is ‘set down’. Within macapat recitation there are a number of
timing techniques used to emphasize the sèlèh note and thus mark the ends of lines. The simplest of
p. 472
these is to slow slightly on the approach to the last two syllables and then hold the penultimate note for
longer than usual, creating tension by delaying the nal pitch, the aural expectation of which has been set
up by the material leading up to it. Although ornamentation is not much used in macapat, it is sometimes
added at this point, making the approach to the nal pitch of a line signi cantly distinct from other
cadential moments, and making the extension of the penultimate pitch more than a simple pause.

Because the recitation of macapat is usually performed by one person at a time, the overall pace, length of C22.P28
pauses between groups of syllables, and approach to ending lines can be chosen and varied based entirely on
the preferences of the singer. While each singer usually maintains a consistent approach for each metre,
when another singer takes over during the recitation of a long text they may interpret the same metre in a
slightly di erent way. Timing is, therefore, an essential element within macapat, ordering the ow of
recitation in a way which marks it out as heightened rather than natural speech, and marking important
points within the textual and melodic structure. The timing is, however, exible because it is an
unaccompanied, solo genre, and involves no coordination with others.

In order to unpick how a macapat is performed when recited individually, I provide an example in the Pucung C22.P29
metre as recited in the manyura mode in the sléndro scale. This is one of the best-known macapat because of
its small size and relatively few variant forms. The text is from the Serat Wedhatama, written by the fourth
Mangkunegaran prince, and falls into the Javanese poetic category of didactic literature: giving advice to a
prince as to how to nd and practise genuine spiritual knowledge:
Here the apostrophes, and the spacing around them, represent signi cant pauses in the onward movement C22.P38
of the melodic line. In my experience, the way in which this pause is treated varies between individuals, and
can consist of a complete break in sound production, or of holding the pitch and syllable that occur just
before the apostrophe for some time before moving on. Although most singers use signi cantly less vibrato
when reciting macapat than they would in other contexts, this held note is also one place where vibrato
might be used to ornament it. The nal two syllables of each line are also slightly separated by a gap or
pause from the preceding syllables to emphasize the cadence to the end of a line, and, if any is used, this is

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
p. 473 where there might be some ornamentation (grègel), shown here in brackets. At the end of the third line,
approaching the pitch low 6, which is a particularly strong cadential point in slendro manyura, the melody as
I have most frequently heard it performed has a very small two-note melisma which is otherwise rare in
macapat. All of these timing techniques for shaping the recitation are performed at a pace which can vary
widely between individual performers but is internally consistent. The way in which these approaches to
pacing and emphasis through timing are learnt is almost completely through observation and gradual
internalization of norms demonstrated by other people, without any explicit direction or instruction. A
great deal of latitude in pacing and timing is accepted when people who have learnt in di erent contexts
recite together, but people may also gradually build a new group consensus if they recite together regularly.

Timing in macapat is, therefore, shaped in response to others in the sense that it is moulded by the C22.P39
performance practices of others while learning and, over time, when reciting regularly with a speci c group,
but it remains inherently exible and personal. There are social and musical constraints on timing, but they
act over the long term and are built into a singer’s personal memory of a macapat, rather than being
immediate responses to other musicians’ behaviour during performance.

Palaran: Reframing Macapat C22.S5

While the material is essentially identical, when macapat are used as palaran these rules of timing are put C22.P40
under considerable stress, as the singer balances rhythmic autonomy and creativity with the requirements
of interacting with the more xed structures and hierarchies of the instrumental ensemble. The
transformation from macapat to palaran is made in two ways. First, the vocal line is performed with much
greater exibility and virtuosity, increasing the complexity and range of the melodic contours used to move
between key pitches, and allowing space for large amounts of ornamentation. Second, the vocal line is now
performed in the context of a strict rhythmic framework which constrains decisions about timing made by
both the singer and instrumentalists. The collision of these two somewhat contradictory factors of
increased exibility and constraint are at the heart of the internal strain in palaran.

Palaran are not performed as standalone pieces in the same way that macapat are so performed; they always C22.P41
occur as an interpolation into a srepegan. Srepegan are a group of pieces which are part of the
accompaniment for shadow puppet theatre (wayang kulit). There are many local variations and speci c
individual srepegan for particular purposes, but fundamentally there is a srepegan in each of the three modes
within each of the two tuning systems of Javanese gamelan, sléndro and pélog. In sléndro, for example, the
core repertoire of srepegan consists of one for each of the modes or pathet: sléndro nem, sléndro sanga, and
sléndro manyura. Srepegan are part of the complex of generic mood music for wayang accompaniment, and
sit in the middle of a three-level hierarchy of pieces which express increasing dramatic tension through
p. 474 increasing simplicity of melodic line and increasing density of strokes on all instruments (but
particularly the lower-pitched structural instruments). The most expansive and melodically complex genre,
which accompanies scenes such as the arrival of expected guests or the orderly performance of court
protocols, is the ayak-ayakan. This has a kenong stroke on every second note of the main melody and a
kempul stroke on every fourth note:
C22.P42

The most intense genre is the sampak, which accompanies ght scenes or dramatic upheavals, and has a C22.P43
kempul stroke on every note of the main melody and a kenong stroke before and with every note:

C22.P44

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
Srepegan sit between these two extremes: they accompany travelling, unexpected arrivals, and the C22.P45
beginning of con icts, and have a kenong stroke on each of the melody notes and a kempul stroke on every
second note:

C22.P46

This structure continues as the basic accompaniment for palaran but, while in a srepegan the pitches played C22.P47
by the kenong and kempul players are tied to the target pitches of each four-note grouping (the sèlèh), in a
palaran they relate to a more general sense of points where the vocal melody comes to rest in a signi cant
way, whether this is at some point during a line or at the end of one. This regular and constant ow of
pitched strokes continues while the singer draws out and beauti es the melodic line.

A possible rendition of the pucung notated above, when performed as a palaran, is given here with possible C22.P48
pitches for the kenong given above the vocal line:

C22.P49

p. 475 Short vocal interjections called senggakan may be inserted between lines of a palaran by the singers who are C22.P61
not currently singing the palaran. For the sake of contrast, the standard practice is to have male singers
performing senggakan if the soloist is female or vice versa. Senggakan are rhythmically metrical and thus
provide some contrast with the drawn-out exibility of the rest of the palaran. They also provide small
oases of xed rhythmic material where the ensemble can regroup if necessary. Senggakan also act as a
bridge between the nal sèlèh note of the preceding line of melody and the next pitch that the vocal line will
be aiming for. In this way they can also act to remove some of the uncertainty as to which note the kenong
player should choose next.

In the performance of macapat timing is a matter of individual aesthetic choice, with the only constraints C22.P62
being the division of phrases into four syllable units, and the models which each singer has picked up and
developed in the process of learning how to recite. When the same material is used as a palaran, singers are
expected to extend and embellish the melodic line whilst accommodating it to the timing constraints of a
small ensemble of musicians all seeking to respond to each other from moment to moment, applying their
own role’s rules of timing and remaining smoothly coordinated.
Issues of Timing in Palaran: Flexible Melodic Accompaniment and C22.S6

Coordinating Cadences

I now examine two of the speci c issues of timing that arise in the context of this collision of vocal C22.P63
exibility and ensemble structure: the response of the instrumental players to the singer’s position within
the melodic line, and the interplay between singer, drummer, and ensemble when approaching cadences

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
that are strong enough to be marked with a stroke on a gong suwukan or gong ageng.

Flexible Melodic Accompaniment C22.S7

One of the primary timing issues within a palaran is the coordination of changes in pitch in the underlying C22.P64
structural framework. The pitches which are played by the kenong and kempul, and danced around by the
soft instruments, establish the aural destination for the singer and the listener, providing a background
which makes sense of the singer’s melodic and rhythmic ornamentation. The notes being provided are,
however, also important for a singer in orienting themselves, and have the potential to be disruptive if
p. 476 misplaced. The structure of kenong and kempul strokes is such that the kempul player is playing once for
every two strokes on the kenong and chooses which pitch to play based on the previous note played on the
1
kenong. The following example shows how this works, but the rapidity of change in pitch here is for clarity
of demonstration and is not representative of normal practice:

C22.P65

In order to avoid clashes of pitch, the kenong player cannot change note on every second stroke, when it is C22.P67
joined by the kempul, but must wait to change at a time when it is playing on its own, allowing the kempul
player time to adjust before they choose a note. This dramatically reduces the moments at which the kenong
can change pitch in order to accompany the singer, and means that split-second decisions must be taken as
to whether to change slightly early or slightly late.

This decision about when to change pitch is one of the main tension points within the ensemble. If the C22.P68
singer wants to start the next phrase but the kenong player is still playing the previous target pitch, then the
singer will have to wait until the structural instruments have moved before being su ciently oriented to
launch the next phrase, disrupting their overall sense of pacing for the palaran. If the structural instruments
change pitch too early, the vocal melody will arrive at an important pitch without the appropriate support
and necessary sense of cadence, disrupting the overall progression between key pitches which makes up the
contour of the palaran. For these reasons, singers often become frustrated with the musicians playing
kenong and kempul when rehearsing and performing palaran. This tension is clearly felt within the
ensemble, and can cause amusement, concern, or excitement among the majority of the ensemble who are
not playing during palaran. The extent to which this is apparent to observers from outside the ensemble
varies depending on their level of knowledge or connoisseurship, but the fact that at least half of the
musicians are also watching and listening, potentially with bated breath, for the un/successful negotiation
of these pitfalls means that the ensemble can be signi cantly more emotionally transparent during palaran,
and audiences do pick up on this.

The soft instruments weave a fabric of patterns that revolve around and lead towards the fundamental C22.P69
pitches played by the kenong. For the musician playing the gendèr, (a 14-key metallophone played with two
beaters which weaves intricate layers around key melodic ideas), however, their more general role as a
provider of pre-emptive, supportive pitches for singers (thinthingan) means that they are often thinking
about pitch in a di erent way. Experienced gendèr players do not simply elaborate on the target pitches, but
move between giving starting pitches for melodic contours, providing nuanced accompaniment throughout
the line, and then setting up nal target pitches. Not being constrained by the formal structure means that
the gendèr player can follow the timing of the vocal line much more closely.

p. 477 Coordinating Cadences C22.S8

The most crucial point of timing tension in palaran occurs in the approach to the places in a particular C22.P70

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
macapat where the sense of cadence at the end of a line is su cient to merit a stroke on a gong rather than a
kempul. In the pucung given as an example above, there are usually three gong: one at the end of the rst,
third, and fourth lines of each stanza. Many of the longer macapat, however, have four or ve gong per
stanza, and so this particular moment of coordination occurs frequently. The gong stroke has to occur at a
point in the structure when the kempul would normally be played, on the second of a pair of kenong strokes.
The arrival of this point/occurrence is signalled by a speci c pattern of strokes on the ciblon drum which
adds a sense of rhythmic cadence to the existing melodic cadence, and which also speci es exactly when the
time for the gong to be struck has arrived. This changes the location of control within the ensemble for the
duration of the approach to the gong, shifting it from being predominantly with the singer to being with the
drummer. While the singer is performing those parts of the palaran which do not lead to a gong stroke, the
drummer plays patterns which maintain a pulse, generate excitement through the creative use of the
multiple distinct sounds available on the ciblon drum and judicious syncopation, but do not create a strong
sense of cadence or stress. The patterns used to signal a gong, however, generate an extremely strong sense
of pull towards the nal beat. While there are many variations on these patterns, they belong to a series of
similar patterns used to signal arrival at a gong in many other settings. For example, they occur in the kébar
patterns used to accompany dance, or the pattern used to signal a suwukan which marks a strong melodic
cadence in a position where there is no structurally required gong in an inggah. Hearing and responding to
these drum patterns are a core part of learning how to play karawitan well and, as a result, they are hard-
wired into Javanese musicians: players will always play a gong or a note on the instrument they are sitting at
in response to this type of pattern even if a drummer idly plays one during a tea break.

In terms of timing, this means that, once the drummer has started to play the pattern signalling a gong, the C22.P71
kempul player knows exactly how many more strokes they have to play on the kempul before the gong will
replace it, and the singer is suddenly signi cantly more constrained as to how they perform the rest of the
phrase. The singer is now obliged to make their variation and ornamentation of the vocal line t into the
remaining time, so that they land on the nal pitch at the same time as the gong stroke. From an ensemble
focused on the singer’s individual manipulation of time and providing appropriate pitches in a stable and
open-ended rhythmic groove, there is a sudden shift to being an ensemble focused on the drummer and
providing a xed and precise response which, once triggered, is completely in exible, controlling even the
singer. Palaran, therefore, alternate between two di erent approaches to timing, shifting between taking a
lead from the singer or the drummer.

It is the attempt to make these shifts as unobtrusive as possible that leads to many of the potential pitfalls of C22.P72
p. 478 performing palaran. The problems discussed here are a selection from the many ways in which a palaran
can go wrong that I have personally witnessed or participated in. The ingrained and de nite nature of the
drum signal for a gong is so strong that it is almost physically impossible for a well-trained musician to play
2
the gong at any time other than the end of it. This generates strain for the gong player if the singer appears
not to be going to land at the end of the phrase at the point that they know they must play the gong: they are
aware that they are in theory accompanying the singer but it is extremely hard to ‘disobey’ the drum signal.
Therefore, if the singer and the drummer appear not to be in step with each other on the approach to a gong,
the possibility of failure arises, and this can only be averted by the singer adjusting their timing in order to
follow the drummer. Occasionally a drummer may forget that a particular line ends with a gong and so does
not start the appropriate drum signal. This means that the singer has to make a decision as to whether to
sing straight through to the end of the line, writing o the gong, or to pause long enough that the drummer
realizes what they should be doing and starts the signal. This is very hard to do in a way that does not
disrupt and distort the previously established pace and ow of the palaran, and thus make the problem
apparent to people outside the ensemble. The drum signal for a gong, the placement of gongs within each
poetic metre, and the cadential melodic movements that ag up the end of sung phrases, are familiar
enough to any Javanese listener that, even for the non-connoisseur, their successful coalescing at the end of
lines is felt and enjoyed. In trying to overcome these di culties—inappropriate timing changes in the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
supporting pitches provided by the instrumental ensemble, or failures to arrive at the gong simultaneously
—speci c advice is given to singers and drummers in learning and rehearsal contexts, and ongoing advice
may be given to kenong players even during performance.

Teaching Timing Techniques C22.S9

In my experience of being taught to play karawitan by professional Javanese musicians, both at ISI and in C22.P73
groups outside the academy who are predominantly from performing arts families, there is almost no
acknowledgement made that the techniques, knowledge, and musical responsiveness required might in any
way be di cult to acquire. The general assumption is that the highly interrelated and, for them, completely
internalized system of rules are self-evident (and should be so even to a foreign learner!). The one exception
to this that I experienced was when being taught how to sing and play palaran—a phenomenon which
clearly marks this genre as something out of the ordinary. In this particular case I was taught very speci c
and considered ways of behaving, whether as a singer or drummer, which would reduce the likelihood of the
problems discussed earlier occurring. These were also imparted proactively and explicitly, rather than
through the usual techniques of giving an example, listening to my repetition, and then providing advice
only in response to my shortcomings in repeating the phrase. The conversations involved in the teaching
p. 479 and rehearsing of palaran were remarkably similar to those reported by Dueck (2013) in the case of jazz
musicians working out an ending. They involved explicit invocation of shared transferrable knowledge,
derived from common repertoire, previous playing experiences, and well-known recordings, and relied on
assumed knowledge of the responsibilities carried by speci c instruments or vocalists in the performance of
palaran. They also always combined the sharing of knowledge to ensure smooth performance with a great
deal of one-upmanship.

When learning to sing palaran I was expected to thoroughly internalize the syllabic structure of the C22.P74
particular macapat being used, and which lines ended in a gong, in order to make it possible to respond
exibly in performance without deliberation or pausing to think. This was particularly stressed in palaran
based on longer macapat with lines of varying syllabic length. The ability to split a line of text into blocks of
four syllables, or the di erent ways of dividing lines of odd-numbered syllables, and to do this consistently
with di erent texts in the same metre, was repeatedly emphasized as fundamental. As well as being taught
to know precisely where I was in the metric structure of the text, I was also taught to maintain a constant
awareness of the underlying structure created by the alternating of kenong and kempul strokes. One of my
vocal teachers, Pak Darsono, actually encouraged me to copy his technique of maintaining a sense of the
structure when practising singing palaran by hitting his left knee for the kenong strokes and his right knee
for the kempul strokes.

An awareness of where one is in the instrumental structure is fundamental to some of the other techniques C22.P75
used to avoid di culties in palaran. As already discussed, the way the vocal line coincides with the position
of the kenong and kempul is a potential problem. In order to avoid the kenong player feeling that they should
be changing note at a time when they cannot (because the kempul is playing with them or being forced to
stay on an inappropriate note while the singer has already moved on to the next part of the melody, which
would be better supported by a di erent pitch), the singer can deploy speci c approaches to timing. I was
taught to start a new phrase at a point in the structure just after the kenong and kempul coincide. This means
that each phrase of four syllables also then tends to nish at a point just before they coincide, allowing the
kenong player space to shift pitch if necessary. Pausing before beginning a phrase in this way falls within the
normal pattern of pausing before sections of syllables in macapat, and so does not seem contrived or
obtrusive.

The drum signal that cues a gong starts on a kempul stroke. It is in the interests of the singer to have the shift C22.P76

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
between their leading the ow of the palaran and their following and adapting to the drummer occur at a
place which is helpful for them and under their control. Starting phrases after the kempul as previously
described, so that they also nish just before another kempul stroke, means that the singer gives the
drummer an opportunity to start the signal in the break between sections of the line. This then allows the
singer to time the nal phrase approaching the gong precisely, without any need to catch up or extend the
melody. This, of course, relies on the drummer being aware of this and also knowing when in a given line to
start the signal for a gong. When I was being taught to drum palaran by di erent teachers from those who
taught me to sing them, a great deal of emphasis was placed on internalizing the structure of macapat so
p. 480 that I could recognize when there were four syllables left in a line ending in a gong and start the drum
signal at this point, thereby providing an appropriate space for the singer to nish the line and end on the
gong. When learning both how to sing and to drum a palaran, therefore, I was given speci c instruction on
how to be as helpful as possible to the other partner in this negotiation of timing.

Even with these techniques, di culties may still arise in the approach to a gong, and singers demonstrate C22.P77
speci c ways of dealing with them. If a singer has arrived at the place in a line of a palaran when they think
the drummer should be starting the gong signal and this has not happened yet, they frequently pause,
perhaps simply holding the note or ornamenting it extensively to extend it, until the drummer becomes
aware of what is happening and starts the signal. If the drummer starts the gong signal at a time which
means that the singer has more or less time in which to complete the nal phrase than is optimal, then the
singer adapts by altering their interpretation of the melody. If there is less time than expected or desired,
then they will cut back to a less elaborate version of the fundamental melodic contour, essentially reverting
to a macapat style approach, leaving any ornamentation to the nal cadence. If there is more time available
than anticipated, then the singer will extend the phrase by ornamenting the movement between each note
and possibly add more than one set of ornaments to the nal cadence, thus lengthening the phrase in terms
of timing while maintaining the integrity of the fundamental melodic contour. The following examples give
a sense of how this is achieved. The rst example shows a pared-down, unexpanded way to land on the gong
2; here there is enough ornamentation to provide the sense of an important cadence, but not enough to
extend the phrase signi cantly. The second example shows how the same phrase might be expanded to last
signi cantly longer. The larger notes here are the core pitches of the melodic contour, and the smaller notes
are potential ornaments:

C22.P78

These approaches to avoiding problems in coordinating timing are taught explicitly, but still within the C22.P82
con nes of lessons, and singers and drummers are expected to deploy them smoothly and unobtrusively in
performance. There is one response to the di culties of palaran, however, which is sometimes employed
visibly during performance.
Correcting Live in Performance C22.S10

Javanese professional musicians, and non-professionals with some musical training, relish being able to C22.P83
p. 481 demonstrate their knowledge. In a rehearsal or informal performance situation this might involve calling
out note names if any of the players appears to be lost, or vocalizing drumming parts. In the context of a
formal performance, however, they might join in with some element of the ensemble such as the clapping,
shouts, or rhythmic phrases performed by the male chorus (gérong), but it is extremely uncommon for them

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
to actually intervene by correcting or directing individual players. This is another of the places in which the
anomalous nature of palaran becomes apparent, as musicians within and without the ensemble often hold
up ngers to indicate the pitch that they think the kenong player should be playing, whether this is because
they think the kenong player is lost or they disagree with the note that has been chosen. It is not uncommon
for there to be more than one person indicating di erent pitches simultaneously, and this can be seriously
disconcerting if one is playing kenong. I have never experienced a singer indicating desired pitches whilst
singing in Java, but when there is more than one female vocalist present, those not currently singing may
request speci c pitches in the same way.

Conclusions C22.S11

Timing is an essential component of the idiomatic performance of both macapat and palaran. The way the C22.P84
core material is interpreted in the two genres is, however, noticeably di erent, and many of these
di erences are related to timing, particularly the length of time given to each section of the melodic
contour, and the level of constraint on the singer’s interpretation generated by other players being involved.
The exing of vocal virtuosity within a xed ensemble structure creates acknowledged di culties in
coordination, and speci c techniques for overcoming these di culties are used and, unusually, explicitly
taught by Javanese musicians. The rules of timing used when reciting macapat emphasize simplicity and
clarity. However, alongside intimate knowledge of the metre’s syllabic structure and associated melodic
contours, they provide the basis of many of these techniques for avoiding problems of coordination in the
far more elaborate and complicated context of palaran singing. The speci c aesthetic created by splitting
lines into groups of syllables and generating emphasis through pauses and ornamentation a ords singers
the possibility of adjusting where their vocal line sits in relation to the underlying structure at key
moments, thereby giving instrumentalists the opportunity to accompany the vocalists more accurately and
reducing the amount that they have to adapt their singing in order to t in with the rhythmically xed
approach to a gong note.

Even though these measures are taken to reduce the impact of the di culties inherent in palaran, they never C22.P85
completely remove the sense of risk and musical labour. This, I believe, is part of the reason that palaran are
disproportionately popular compared to other elements of karawitan and are also signi cantly more popular
than recitation of macapat. It also explains their use, in theatrical contexts, as a vehicle for combative verbal
exchanges, ranging from light banter to preludes to battle. Many of my Javanese friends who are not
p. 482 musicians associate palaran exclusively with this context, rather than their use in purely instrumental
settings, and experience them as intimately linked to the production of dramatic tension. The clear sense of
musicians responding and adjusting to each other in real time, and the possibility of things going wrong,
are part of what makes this particular genre so alluring and what links it to the expression of tension and
excitement.

The example of macapat and palaran presented here is a very speci c, culturally bounded case, with timing C22.P86
being used in highly idiomatic ways to emphasize poetic structure, highlight cadences, and beautify melody.
Despite its speci city, however, this case study shows how the necessity to coordinate timing can lead to
approaches to the transmission of skills and knowledge, and in-performance negotiation and adjustment of
control, which are similar to those deployed in other genres where musicians lead and follow each other
without a xed script. The fact that the core material for both genres is very similar provides a particularly
sharp focus on both the extra tensions in coordinating timing that arise when interpreting a sung text with
multiple musicians rather than a single singer, and the potential for creativity and excitement that this can
generate.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471104 by National Science & Technology Library user on 26 May 2023
Notes
1. The actual pitch played by the kempul player may not be exactly the same as that chosen by the kenong player. The modal C22.N1
rules governing this are too complex to discuss here and are not strictly relevant to issues of timing.

2. This sense of following the drum pattern to the gong involves synchronizing oneʼs internal pulse to the drummer, but only C22.N2
for a specific moment. As such it does not involve entrainment in the sense described by Clayton (2012), which involves
ongoing interaction between two pulses. It certainly represents a form of embodied knowledge based on appropriate and
timely response to signals, which is also linked heavily with social and moral responsiveness in Javanese discourse
(Roberts 2015: 16–17, 27–28).

References C22.S12

Arps, B. (1992). Tembang in two traditions: Performance and interpretation of Javanese literature. School of Oriental and African C22.P87
Studies, University of London.
Google Scholar Google Preview WorldCat COPAC

Brinner, B. E. (1995). Knowing music, making music: Javanese gamelan and the theory of musical competence and interaction. C22.P88
University of Chicago Press.
Google Scholar Google Preview WorldCat COPAC

Clayton, M. (2012). What is entrainment? Definition and applications in musical research. Empirical Musicology Review 7(1–2): C22.P89
49–56. https://doi.org/10.18061/1811/52979
Google Scholar WorldCat

Darnawi, S. (1982). A brief survey of Javanese poetics. PT Balai Pustaka (Persero). C22.P90
Google Scholar Google Preview WorldCat COPAC

Do man, M. (2011). Jamminʼ an ending: Creativity, knowledge, and conduct among jazz musicians. Twentieth-Century Music C22.P91
8(2): 203–225. https://doi.org/10.1017/S1478572212000084
Google Scholar WorldCat

Dueck, B. (2013). Jazz endings, aesthetic discourse, and musical publics. Black Music Research Journal 33(1): 91–115. C22.P92
https://doi.org/10.5406/blacmusiresej.33.1.0091
Google Scholar WorldCat

p. 483 Roberts, J. (2015). The politics of participation: An ethnography of gamelan associations in Surakarta, central Java. Doctoral C22.P93
dissertation, University of Oxford. https://ora.ox.ac.uk/objects/uuid:c8975102-b7c8-4e07-874d-9bd3371de216
Google Scholar Google Preview WorldCat COPAC

Saprodjo, S. G. (2002). Primbon cakepan tembang lengkap: Klasik populer dan kreasi baru untuk bawa, gerong, sindhen, dolanan, C22.P94
p. 484 langgam. Cendrawasih.
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
CHAPTER

23 Here at the Bottom of C23


the Sky…: Negotiating Time
through Phrase, Form, and Tradition within a New York
Performance Network 
Nathan C. Bakkum

https://doi.org/10.1093/oxfordhb/9780190947279.013.26 Pages 485–504


Published: 08 December 2021

Abstract
While improvisational interaction is often understood as the momentary give-and-take between
musicians in performance, this chapter considers ways in which relationships between improvisers
function at a range of di erent timescales. Organized around extensive conversation with drummer
Tom Rainey, bassist Drew Gress, and other members of their performance network, the chapter reveals
ways in which this group of musicians employs shared cognitive schemas, collaborative approaches to
composition, and common musical histories to shape small- and large-scale musical time at the levels
of gesture, form, and tradition. Through analysis of musical examples from several of Rainey’s
working ensembles, the chapter demonstrates speci c ways in which those shared conceptions of time
are realized in sound.

Keywords: Tom Rainey, Drew Gress, jazz improvisation, interaction, cognition, form, tradition
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

In the band with [saxophonist] Tony [Malaby] and [keyboardist] Angie [Sánchez], some of my C23.P1
favorite moments are when we hit these really simple things. I might just be playing quarter notes
for a while. Nothing but quarter notes on a cymbal. But the way it’s all tting together, I could just
keep doing it forever. And in a weird way, no matter what I’m doing, no matter how busy or
abstract or loud or soft or anything is, I kinda really want it to be this feeling of playing quarter
notes on a cymbal. I want it just to have that ease and that clarity, even if it’s incredibly dense.
There’s also times when I could be playing duo with Tony or with [saxophonist] Tim [Berne] and I
know we’re in a tempo, but I don’t know what it is. It’s just hooked up. And it might be three
di erent tempos, but it doesn’t really matter.

1
Tom Rainey C23.P2
FOR percussionist Tom Rainey and his bandmates, time—groove, phrasing, form, and tradition—is actively C23.P3
p. 486 negotiated through intense, sustained improvisational interaction. This is an embodied and distributed
practice, re ned by individual musicians and ensembles through ongoing study and steady communication
as they develop individual gestural vocabularies while attuning to collaborators’ tendencies and
expectations. These embodied practices extend beyond individuals and discrete ensembles, toward larger
performance networks of musicians with overlapping relationships and shared practices. Rainey and his
many collaborators have developed robust interactive models that allow them to negotiate the performance

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
of time at multiple timescales simultaneously, in ways that both reinforce and expand the range of expected
interactive gestures and phrases, collective forms and performances, and shared histories and lineages
within the performance network. Ultimately, a musical community and economy coalesce around these
shared understandings and collective performances of musical time. While their choices are constrained by
genre and tradition, shared experiences, and the needs of speci c compositions, these musicians
continually explore communication patterns that might resonate within the collective in new ways. Rainey
says of his primary performing mode:

It’s really like moment-to-moment intense concentration to try to just make the music sound the C23.P4
best you can. Playing with Tim or playing with [bassist] Drew [Gress], or Malaby or whoever, if I
just started falling back on ‘okay, this works,’ they wouldn’t dig it. But it requires a lot more of
yourself to do this. You have to really give up, as best you can, your ego and whatever thoughts of
personal gain, and just try to make the music sound as good as possible.

(Chinen n.d.)

Whether the music requires quarter notes on a cymbal, a straightforward rock backbeat, or a dense barrage C23.P5
of sound, individual contributions are judged collectively and compositionally. Together, the musicians in
Rainey’s performance network shape sound, experience, and time itself.

In this chapter, I explore several di erent layers of interaction that intersect in the workings of this network C23.P6
of musicians, focusing on the ways through which their performances collectively shape and are shaped by
shared temporal understandings at the levels of gesture, composition, and the broader traditions of the
music. Although interaction is typically viewed as belonging to the immediate and momentary, this chapter
highlights how interaction is also built into successively larger timescales. Beginning from a consideration
of the cognitive and interpersonal foundations of improvisational interaction, the chapter moves toward an
analysis of ways in which Rainey and his collaborators collectively generate larger musical forms and assert
their musical identities in relation to a range of American musical traditions.
p. 487
Making time in the moment: cognition, gesture, interaction C23.S1

The collective creation of musical time begins at the level of gesture and phrase with the momentary back- C23.P7
and-forth of musicians in dialogue. For musicians to develop systems of interaction, and for them to learn
to work successfully together, they must engage in the ‘moment-to-moment intense concentration’ Rainey
describes, and they must commit to a exible and uid ethic of musical compromise and adjustment. For

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
each individual, the challenge begins as a juggling act between ful lling the needs of the composition,
responding to bandmates, and contributing meaningfully to the improvised dialogue. Rainey says, ‘I’m
trying to improvise, and I’m trying to feel like I have a part as much as everybody else has a part, even
though it’s not written.’ In discussing Open Loose, a working group in which he performs with Tony Malaby
and bassist Mark Helias, Rainey characterizes his role slightly di erently, saying, ‘It’s all reaction […] that’s
all I’m really trying to do. I’m just trying to sustain the groove, but also catch bits of Tony, catch bits of
Mark, whatever pops out at me, and acknowledge that in some way, even if it’s in a contrary way, but some
sort of acknowledgment of what they’ve played.’ Addressing the ways musicians must adapt to di erent
composers and bandmates, Drew Gress simply says, ‘You don’t speak with the same voice or say the same
2
things to everyone you meet; you have to be speci c to the situation.’

As musicians engage one another in the real-time organization of musical time through the creation and C23.P8
manipulation of rhythmic, harmonic, melodic, timbral, textural, and formal systems, they develop
individual and collective mental models and embodied strategies. Tom Rainey says, ‘Ultimately, no matter
what I’m playing, I want it to get to the point where the brain is really not involved—at least consciously.
The only thought I’m having is like “WHOOO!” ’ Elsewhere, he says, ‘If my brain is involved too much then
it’s not going well. If I need to be thinking about what I should be doing or how I should be doing it or if it’s
going well—if there’s a lot of thought going on, something’s wrong. And something goes wrong quite
often.’

In these comments, Rainey equates the brain with conscious thinking and intellect. Nonetheless, in C23.P9
qualifying his ideal performing experience as one in which he is not consciously focused on his performative
choices, he reveals an understanding of his own cognitive system as not solely under the jurisdiction of the
brain. This embeddedness and interconnectivity of brain function, referred to as ‘enactive cognition’ by
neuroscientist Francisco Varela (Varela et al. 1991), ethnomusicologist David Borgo (2006), and educational
theorist Wayne Bowman (2004), asserts ‘mind’ as a distributed system encompassing brain, body, and
environment. Bowman summarizes:

To the extent that higher-level cognition and lower-level motor control share basic neurological C23.P10
mechanisms with each other, ‘mind’ is a profoundly distributed entity: the mind is not in the
p. 488 brain, but in the vast network of neural interconnections that extend throughout the body. In
this way, the body is in the mind. Mind is rendered possible by bodily sensations and actions, from
whose patterns it emerges and upon which it relies for whatever intellectual prowess it can claim.
At the same time, the mind is in the body, in the sense that mind is coextensive with the body’s
neural pathways and the cognitive templates they comprise.

Further still, to the extent that these neural pathways and cognitive schemata arise from a body’s C23.P11
interaction with an experience-shaping environment, mind extends beyond the physical body into
the social and cultural environments that exert major in uence on the body and shape all human
3
experience. Thus, both body and culture are implicated in and constitutive of mind. (2004: 36)

Similar to the way that such theories of enactive mind suggest a model of distributed interaction within C23.P12
individuals and across collectives, so do improvising musicians like Rainey articulate an interconnectedness
of body, brain, and collective that most often manifests itself as a feeling of automatic performance, bodily
4
knowledge, and ow. Audiences might recognize it as a deep shared connection between performers
involved in simultaneous improvisation.

In Rainey’s view, he is best equipped to engage with his bandmates at the moments in which his cognitive C23.P13
system is most broadly distributed.

I’m not an intellectual person […] but I do have a brain and I have a certain intellect and that’s part C23.P14

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
of me and I want that to inform the music up to a point. But I guess I don’t want it to be engaged
while I’m actually doing it. I think I want all of the thinking I’ve done about the music and
everything to sort of organically ooze out of me somehow. And that’s when it feels best, when the
arms are doing it. I’m not directing anything. It’s just happening, and it’s like an immediate
response to whatever I’m hearing and it happens very naturally.

Jazz scholars have sometimes struggled to account for the choices made by individual improvisers in the C23.P15
midst of performance. A move toward understanding the cognitive processes of jazz performance as
distributed, bodily, and communally grounded in time allows us to understand these choices as re ective of
the social system within which they are performed and as part of a broader jazz tradition.

Drew Gress describes a wide variety of collaborative interactive frameworks as central to his own C23.P16
improvisational organization of musical time. Tellingly, each of Gress’s descriptions focuses on a speci c
interactive, rather than soloistic, logic. He says, ‘Sometimes, I like to experiment with playing something
that is compositional within my own sphere. One way to relate is to have whatever it is you do have its own
integrity so it holds together as itself. It just happens to coexist in time with what the other musicians are
playing.’ While he refers to the compositional integrity of his own performance, he understands musical
interest as generated through the layering of multiple parts, gestures, and timeframes. In a di erent
situation, he says, ‘Another [method] may be having this particular piece be about some kind of physical
challenge you set yourself on the instrument and try to get to a place where you’re pushing your limits to
p. 489 see what kind of energy that creates.’ Here, the goal of establishing a physical challenge is not to push
oneself as an individual, but rather to create a speci c energy with which other musicians may engage.
Finally, he says:

Sometimes you just take a passive role and just kind of reconcile what someone else is doing. Other C23.P17
times you start to dictate things. Sometimes you’ll choose to pile on something that’s already
crowded, and other times just totally lay out. Sometimes, you might play in the same vibe as
someone else, but the focus might be totally on rhythm and you’re unconcerned about pitch just to
see what that creates. Like what if I totally turn o my usual harmonic thing, whatever that is, as
far as, OK, I hear the next note…will I play that note or not? Maybe I hear the note but don’t play it
and play something that’s intentionally not that note. So it’s really about testing the waters with
somebody, just to see what the response might be.

Throughout this description, Gress thinks compositionally, attempting to create interesting time/spaces of C23.P18
density and texture by alternately reconciling other musicians’ choices and introducing new elements to the
conversation. He summarizes his improvisational philosophy by saying, ‘I think it’s basically about
strategies of how to create di erent energies as an element of variety and seeing what that provokes in the
other players.’ Even within well-established interactive systems, Gress and his bandmates continue to push
each other into uncomfortable, unfamiliar spaces and temporalities.
Composing together in time C23.S2

As musicians piece together gestures, phrasings, and textures, larger structures of musical time begin to C23.P19
emerge from their collective play. Moving away from work at the micro-level and the immediacies of groove
and phrasing toward broader timescales, the musicians in Rainey’s performance network pay substantial
attention to the architecture of complete songs, performances, and recordings. Referring to the work of the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
fully improvised Paraphrase trio, featuring Gress and saxophonist Tim Berne, Rainey says, ‘Everybody to
varying degrees is a composer, so I think everybody, no matter what the group is, is conscious of that big
picture.’

When engaging with new compositions, Rainey acknowledges that he and his collaborators must take their C23.P20
time in bringing that big picture into focus, as their temporal focus moves toward building satisfying
musical structures over multiple rehearsals or performances. As Rainey says, ‘I really get into a problem of
trying to run before I can walk with music a lot of times. I want it to immediately be inside me and feel free
over it. And it just doesn’t work that way. So there’s times when it’s really appropriate for my brain to be
involved and say “Pay attention, look at the music, don’t worry about being creative.” ’ But as
p. 490 compositional familiarity increases, musicians’ understandings of the pieces inevitably colour their
performances, fundamentally a ecting the shape of the composition itself.

As this group of musicians begins shaping larger interactive systems around particular pieces or ensembles, C23.P21
some ideas quickly take hold, and others fail to resonate. Rainey says, ‘If you’re on tour, and let’s say you’re
doing 15 gigs in two-and-a-half weeks, then there’s a lot more liberty to try di erent stu . Because if it
doesn’t work, it’s no big deal—you’re going to do it again tomorrow night, and you can try something else
or go back to what was working.’ Describing Paraphrase, Drew Gress concurs, ‘Sometimes it can be
terrifying if we’re having an o night, which we do occasionally. We’ll keep trying, but it might be that
everyone just keeps missing and it isn’t hooked up, but the commitment to improvising remains, and in a
way it’s reassuring to have an awful night every once in a while, because that means you are improvising, I
guess.’

On these ‘o ’ nights, as on the nights when everything clicks and everyone on the bandstand appears as if C23.P22
of one mind, the musicians maintain responsibility for reproducing the interactive structures that frame the
group identity. Greg Urban suggests that without directed action from knowledgeable actors, such
structures simply cannot coalesce (2001: 15–32). Urban focuses on the motion of cultural objects as
foundational to the emergence of social structure. Within jazz communities, these cultural objects might
include compositions; rhythmic, formal, timbral, or harmonic ideas; recordings; standards of
communication; or other communicative devices. As speci c routes take shape, Urban argues that these
‘inertial pathways’ might solidify into shared understandings and expectations.

Once individuals gain con dence and familiarity with one another, musicians remain free to experiment C23.P23
with boundary-testing, recon guration of assumptions, and other types of ‘play’ within and around the
structural frame. Urban refers to these highly contingent and unstable challenges to communal standards as
‘accelerative’ pathways, suggesting that structural change results from the gradual solidi cation and
sanctioning of a new approach, through which an accelerative gesture becomes an inertial one.

As Gress states above, performers generate much meaning and musical interest through just this sort of C23.P24
‘play’ within and against entrenched structures. ‘It’s really about testing the waters with somebody,’ he
says, ‘just to see what the response might be.’ Though Rainey describes much of his working method as
resulting from restlessness, sel shness, and boredom, his commitment to exploration allows him to
continually ‘test the waters’ and challenge other musicians. While many of his choices only momentarily
a ect his bandmates, Rainey’s constant reaching, pushing, and ‘tuning’ gives him ample opportunity to
instigate more sturdy stylistic change to the ways that his bandmates approach full compositions.
Gress marks Rainey’s stylistic variety as one of the drummer’s de ning sonic signatures: ‘I wish it was C23.P25
possible to include a variety of takes for each composition on the [7 Black Butter ies (2005)] CD, so the
listener can hear how di erently he plays from take to take. He’s fearless.’ Rainey recognizes this exibility
as central to his improvisational philosophy, as well, but is typically understated about its signi cance:

p. 491 I’m not really trying to change the stu radically every time, but it might just be a sound thing, I’ll C23.P26
play the written material using di erent sounds, but maybe rhythmically playing pretty much the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
same stu . Or dynamically, maybe suddenly I’ll really bring something down that’s been loud. It’s
those kinds of changes a lot. It’s not like I have this militant agenda about, like, every time it has to
be like it’s never been played before. I’m trying to improvise, and I’m trying to feel like I have a
part as much as everybody else has a part, even though it’s not written. For the written stu , I
don’t want it to sound like everybody’s playing this composition, and I’m just over there fucking
around. I really want to have the feeling that I’m nding my part. And the more times you do that,
of course, you become free to nd other parts.

As a simple example of Rainey’s philosophy in action, consider the two recordings of the tune ‘Huevos’ that C23.P27
Rainey has made with Berne. On the rst, taken from Berne’s 2002 studio album Science friction, Rainey sets
up a composite rhythm with guitarist Marc Ducret that allows the whirling melody of Craig Taborn’s
keyboard and Ducret’s second, overdubbed guitar line to lie comfortably in a mid-tempo pocket. Rainey’s
simple but (relatively) consistent ride cymbal pattern, despite the freedom with which he implements it,
grounds the performance in a 1970s electric Miles Davis-style groove.

By contrast, on Berne’s 2004 live recording of the same piece, Rainey abandons the ride cymbal entirely, C23.P28
moving his most consistent pulse to the hi-hat. Rainey adopts a considerably more exible approach here,
moving freely around the drum kit and focusing on complementing and commenting on the saxophone and
keyboard melody. He ful ls his responsibility to the ensemble in both cases, o ering a continuous
propulsive stream of eighth notes and helping the other players to lock into the uneven melodic cycle.
However, Rainey’s overall approach to the live recording allows Taborn and Berne the opportunity to take
more rhythmic risks, and as the piece continues, the group’s rhythmic concept fractures and recombines
into several di erent types of groove, ranging from the steady 5/4 metre associated with the studio
recording to non-metric, loosely pulsed sequences and entirely unmetred sections. In just this way, over the
course of multiple takes over multiple nights, Rainey and his bandmates might recon gure one another’s
understandings of how they collaboratively shape a particular piece.

Gress sees this kind of experimentation as foundational to his and Rainey’s shared aesthetics: ‘That, I think, C23.P29
is the basis for a lot of the playing we do together—seeing how far we can push. It’s almost like tting a
sleeping bag into those miniaturized containers in a way; it’ll just take the shape of the container (or the
song), but let’s just see how much u y material you can put in there. You just think of the songs as
containers.’ Speaking as a composer, Gress goes on to mark this kind of interactive stretching as a central
goal of his ensembles and a primary consideration in his personnel choices:

One of the reasons I have the people I have in my band is that they’re going to push, bang on the C23.P30
cage, and test the limits of the container. It’s not that [Rainey will] do something drastically
di erent, but within the context of takes on a record, I wish I could [release multiple versions],
because especially in long improvisational sections, the character can be markedly di erent. It’s
p. 492 just all the detail, really. He’s just not on autopilot ever. He’s always right there—an active force
to be dealt with. He never goes to sleep.

Through his choices as a bandleader, Gress acts as a structural caretaker. He cultivates the development of C23.P31
speci c types of accelerative pathways through his choice of performers and the freedom he grants them,
while carefully controlling certain inertial pathways through his compositional choices. Like many
composers in his community, Gress sets up the conditions within which a exible and uid musical system
can ourish.

As a composer, Gress often tries to create spaces in which the shared backgrounds and musical a nity of C23.P32
his bandmates might help to shape his compositions dynamically in performance. While his 7 Black
Butter ies ensemble reveals strong sonic and performative links to the bebop tradition, it also shows the
in uence of the extended techniques of Charles Mingus and members of the AACM, as well as the textural

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
ideals associated with modern classical composition and rhythmic elements of 1970s mainstream rock.
Gress takes an active role in shaping the group’s work through his compositions:

I try to come up with durable material […] So, there’s some aspect of it that I think the improviser C23.P33
is going to enjoy manipulating, and then I will try to sit with the material for a while and see what’s
suggesting itself as the strongest area for improvisation or maybe there’s some inherent structure
with the tune that I think will be more illuminating than if we just play the thing straight through
and just start improvising on the form of the tune.

On ‘Zaftig’, from the group’s self-titled 2005 album, Gress blends cyclic forms with through-composed C23.P34
sections while also leaving space for the ensemble to shape each performance collectively. The composition
features three large sections in unrelated tempos, with the piece opening as a ballad, working toward a mid-
tempo vamp, and returning to a slower tempo for the conclusion. Faced with the di culty of crafting a
compositional solution that would e ectively transition between these distinct tempos, Gress ultimately
called for an improvised answer to this challenge. Rainey says, ‘That took a while, actually, to make that
work. [Gress] had the idea, but he kept trying to do it mathematically. Finally, we were just like “Why
doesn’t [trumpeter Ralph Alessi] just come in with a di erent tempo and not worry about it?” (laughs) But
anyway, we solved it.’ Rainey’s comment simpli es the dynamics of the moment, as the ensemble works
together to realize Gress’s compositional ideal.

After opening with a fully composed ensemble section, a piano solo by keyboardist Craig Taborn serves as an C23.P35
introduction to a four-and-a-half minute collective improvisation, eventually leading to Alessi’s entrance
in the new tempo. Throughout this improvisation, the ensemble creates shifting layers of di ering densities
through their choices, joining together in small interactive cells before splitting and recombining in new
combinations, as outlined in Figure 23.1. After a single cycle through Gress’s prescribed solo form, Rainey
p. 493 breaks away from the collective (represented by the light grey shading in Figure 23.1) and moves through
a faster tempo into a free, phrasal accompaniment, primarily on cymbals. Gress also begins to play more
rhythmically free, allowing Taborn the freedom to establish a faster vamp and transition smoothly into the
saxophone solo. Gress and Rainey reconnect with Taborn’s new tempo, and all three accompany Berne’s
spindly saxophone improvisation.

Just before the 5:00 min. mark, Rainey assertively moves into yet another faster tempo (marked by the dark C23.P36
grey shading in Figure 23.1), and one by one, the ensemble’s other players join him. Taborn and Gress cycle
through a composed vamp at this new tempo, cueing Alessi’s slower entrance. Rainey says, ‘It’s like a
cross-fade of two ideas. Drew’s playing a part, Craig’s playing a part, I’m free but I’m relating to that part.
And when Ralph comes in, it gets real messy.’
Figure 23.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
Interaction and navigation in ʻZa igʼ. C23.F1

Gress provides the ensemble with a series of signposts, but he allows them to choose their own adventure C23.P37
p. 494 through this segment of the piece. The tempo transitions, which are requirements of the composition,
become a collective e ort and an opportunity for the band to stretch out in seemingly in nite directions,
provided that everyone is thinking compositionally.

While ‘Zaftig’ presents a thoughtful collective solution to a complex compositional problem, Rainey’s C23.P38
approach here is consistent with the majority of his work. In each of the working groups with which he
performs, the composers and bandleaders allow the drummer signi cant freedom, making Rainey a nexus
of uidity and change within the community. None of the bandleaders compose speci c parts for him,
instead allowing him to gradually develop his own contributions to the compositions. Rainey says, ‘They
really kind of give me pretty free rein to nd it, and they’re very nicely patient about not getting too worried
about if it’s not sounding great right away and giving me time to gure out a part.’ Describing his process of
working through a typical composition by Helias, Rainey says:

A lot of times he doesn’t even have to tell me, I can just tell by looking at it what it’s going to be C23.P39
like. So once you kind of get an idea for the feel, then I’m just trying to wind myself around these
lines. And maybe I want to play more with the sax in this part than the bass, and […] a lot of times I
just start o really looking at the bass part, at least as I get to know it, so we get that kind of
foundation. And then if he has ideas that he wants me to lay out here or play more with this there
or whatever then we talk about it.

But the process does not end there; as Rainey says, once he nds his part, he turns his attention to ‘ nding C23.P40
other parts’. Gress considers Rainey’s process as a compositional bene t, saying:

I think it’s better in general, with players I respect like Tom, I have the chance to be surprised, and C23.P41
have him make the tune far better than I could’ve imagined it. So I at least want to see what he
comes up with to see if that’s some way that the tune can go. And because he is so compositional,
he can impose a variety on the proceedings no matter what. I just don’t want to restrict him. I want
him to feel free to react to what’s going on.

Rainey’s bandmates regularly re ect on his constant adaptation of parts and reinterpretation of C23.P42
compositions, and they mention how completely they trust him to craft his improvisations in ways that will
enhance the collective performance. Why do composers give such freedom to the player they most associate

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
with restlessness, exibility, and consistent change? It seems the answer lies in the community’s
commitment to the element of surprise; to the uidity of their interaction; to a focus on a speci c type of
process and exploration as de ning and directing their work. As Rainey says, ‘I’m really aware of the fact
that I don’t want to bore these great musicians I play with. I want to kind of keep them surprised, too. I
guess that’s the motivation for it as much as anything.’

This compositional ethic extends to fully improvised settings, as well. Within the Paraphrase trio, Rainey, C23.P43
p. 495 Gress, and Berne have the opportunity to collectively explore a breadth of styles in an open and extended
format. Although Paraphrase is the only one of Rainey’s steady ensembles that works in a completely
improvised idiom, the group members focus on a compositional sensibility informing their work as they
explore a range of musical styles and shared touchstones. Gress says:

[Rainey] approaches his playing in those groups as a composer. So he’s thinking of the piece as a C23.P44
whole. Or the evening as a whole, or the set as a whole. If we’re playing as Paraphrase, then
whatever we’re doing we’re doing in the context of the set as a complete listening experience for
the listener as well as ourselves. So that’s the main consideration, I think—trying to do something
that has compositional elements to it. So in a way, when we’re playing open, my focus tends to be
to impose some kind of order on the proceedings, just so they can kind of be intelligible to the
listener. You’re looking for structure—spontaneous structure.

Separately, Rainey o ers an example: C23.P45

Even though Paraphrase exists within Drew’s band [7 Black Butter ies], even if we’re playing just C23.P46
the three of us in Drew’s band, it’s a completely di erent feeling than it is when we’re playing as
Paraphrase. Which I like. I like the fact that we don’t, whenever we have a trio section in Drew’s
band, it just doesn’t turn into a Paraphrase gig. I think that says a lot about the sensitivity of the
musicians dealing with the di erent situation of the composition versus the open improvisation.

I think what I’m saying is just how it feels to play with Tim and Drew on a Paraphrase gig and how C23.P47
it feels to play with Tim and Drew on a Drew gig is just a completely di erent feeling. I think it’s
obviously the in uence of the compositions and the fact that we just don’t go ‘Yeah, OK, it’s the
three of us playing, so we’ll just get into our Paraphrase thing.’

The group performs set-length pieces, generally working through long-form improvisations between 20 C23.P48
and 45 minutes in length. Describing the trio’s overall philosophy, Rainey o ers a lmic metaphor: ‘It isn’t
just these burning moments—it’s like a little movie or something. I think we all really want that feeling,
that after this piece or pieces are done in a set, that it was somehow a cinematic experience. That there were
these di erent emotional things that happened and di erent scenes that happened and di erent shadings.’

Like Rainey’s other ensembles, Paraphrase moves between unmetred and more groove-based styles, C23.P49
though they work in much longer and more gradual timeframes. On ‘Poetic license’, from their 1997
recording Visitation rites, the group works through four large improvised spaces over the piece’s 20 minutes,
as shown in Figure 23.2. Each of the main improvisational spaces lasts between 2 and 5 minutes, and each
has a contrasting character from the others. In all but one case, the group moves between areas with a brief
(30 seconds to one minute) transitional section, including an intro and outro. The consistency with which
p. 496 the group collectively generates this form gives the piece both a constant motion and a comforting
regularity, bringing Rainey’s cinematic metaphor to life.

Figure 23.2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
Space and movement in ʻPoetic Licenseʼ. C23.F2

In the one case in which the group does not make use of a gradual transition between sections, they rely on C23.P50
another device for creating contrast. Describing one of the ensemble’s standard practices, Rainey says,
‘Something that happens with Paraphrase is that […] it’s really abstract and there’s no discernible tempo to
it at all, and then—it might be my decision or it might be anybody’s decision—and suddenly it’s there,
there’s something being implied’ (2006). Approximately twelve minutes into the performance of ‘Poetic
license’, Berne brings his saxophone solo to a quick conclusion as Rainey and Gress nearly simultaneously
pull their scattered, unmetred playing into alignment. Seemingly out of nowhere, a burbling, aggressive free
improvisation becomes a mid-tempo, funk-in ected bass solo. Whether understood as a ashback, a jump-
cut, or a surprise ‘twist’, this moment disrupts the ow of the piece and begins its long, cinematic
conclusion. It is a moment of collective improvisational drama that can only be brought into being if each
musician commits to thinking compositionally.

Performing lineage and a inity over time C23.S3

While Rainey and his collaborators foreground notions of innovation, uid change, and ‘newness’, and C23.P51
while each group’s sounds certainly bear individual and recognizable sonic signatures, all of the ensembles
considered here also make their stylistic in uences readily known. This reveals a further, more di use
p. 497 timescale at which he and his collaborators are in active negotiation: that of the jazz tradition as a whole.
Within their admittedly small marketplace, the success and continued viability of these groups depends on
their ability to communicate to their audiences a particular historical narrative and their position within it.
The story the players choose to tell through their music o ers a unique and personal presentation of the
jazz narrative and an attempt to contribute to that narrative’s course.

For many improvising musicians, this process of personal and collective lineage-making serves several C23.P52
important purposes both within jazz culture and in communicating cultural identities to broader societal
groups. By connecting themselves to a well-worn and broadly accepted lineage of jazz musicians,
improvisers assert strong connections to the centre of jazz culture. By emphasizing a distinctive selection of
mentors as central to their development, musicians stake out personal space within the community. These
lineage-making processes simultaneously connect musicians to broader communities of performers and
provide an important opportunity for musicians to distinguish themselves within a crowded marketplace.

A shared history, however imaginary or virtual, greatly improves the chances of a successful collaboration, C23.P53
allowing them to negotiate the shorter timescales of individual performances with con dence in their
interpretations of their collaborators’ work. As Gress recounts in discussing his rst meeting with Rainey,
‘We realized we were into a lot of the same music, even though we never played together. He’s two years

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
older than me, but we loved a lot of the same music, and I think that just naturally creates some kind of bond
that you hear things the same way.’

This shared history provides much more than a musical a nity; it gives musicians a shared storehouse of C23.P54
styles and ideas with which to experiment in the moment. Though many musicians involved in free
improvisation intentionally avoid musical gestures that might imply tonal harmony or metric regularity,
Gress, Rainey, and their frequent collaborators generally embrace a variety of styles as vehicles for play.
Discussing his willingness to lock in to a steady groove, Rainey says:

Maybe it simply has to do with the in uence of all the music I like, and all the music I like is a really C23.P55
broad range of stu . Drumming wise, from the absolute most simple rock drumming all the way
through the most abstract, complex, and everything. And that’s all stu that’s in my world. So I
want to be involved in music where I can—without doing exactly these things—have them inform
the music, even if it’s completely open improvisation.

Rainey’s broad stylistic palette, which is shared widely across his performance network, is here simply C23.P56
equated to the speci cs of personal history; he plays music that means something to him, and a variety of
musical styles—some of which have a jazz pedigree and others of which are far removed from that tradition
—have been and continue to be important to him. These stylistic building blocks are united under a
common process shared by many members of the community.

p. 498 Within a community of musicians that prizes stylistic breadth in open improvisation and new composition, C23.P57
Rainey has chosen to draw a direct connection to the core of the jazz tradition in part through repertoire,
building an ensemble around the Great American Songbook. His Obbligato quintet—featuring Alessi on
trumpet, Gress on bass, saxophonist Ingrid Laubrock, and pianist Kris Davis—has dedicated itself to the
exploration of some of the best-known and most widely interpreted jazz standards. His stated process for
choosing the repertoire that group has recorded for its two albums is simple. ‘The most important criteria
was [sic] that everybody knew the song, they weren’t things that people needed lead sheets for. Because of
that, people could be really free with the material’ (Laskey 2017).

With such familiarity and freedom, the group is able to focus its work on the collective creation of large- C23.P58
scale form in a way that draws from a wide array of traditions and ideologies. Rainey articulates the group’s
approach—and, by extension, his larger improvisational philosophy—as deeply connected to the roots of
jazz: ‘It’s not like it’s a new idea. Collective improvisation on a tune—that’s what dixieland’s about. But
then something got lost along the way, as jazz became more of a soloistic art: the obbligato-ing’ (2014).
While the members of Obbligato hold tight to the melodic and harmonic materials that de ne each
composition, they dance around the edges of those foundations—playing with layers of focus and
abstraction to reveal each tune gradually. As Kevin Whitehead writes in the liner notes to Obbligato’s self-
titled 2014 album, ‘Melodic variation is the heart of their method—they acknowledge the changes but are
very much playing the tune not just the frame, making ve-way counterpoint coherent. Think of the
collective improvisations as multiple obbligatos, responses to a silent theme’ (2014). Rainey articulates his
expectations of the group:
The one thing that I asked of everybody is to just deal with the form and keep the form. You can C23.P59
take it as far a eld as you want, but you have to be able to snap back to it at any point. On the
record, we really adhere to the forms. If you were to listen closely, all of the forms are very much in
tact, no matter what is going on on top of them. Like I said, it’s a playground, and so you can play
with it in any way that you want. You just need to know where you are in the playground so you can
join up with the others.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
(Laskey 2017)

Through this ethic of focused listening and collective improvisation, the members of Obbligato spin a series C23.P60
of time-worn melodic and harmonic forms into fresh and surprising shapes.

Obbligato’s 2014 album includes a recording of Jerome Kern’s 1933 composition, ‘Yesterdays’. Over the C23.P61
course of the recording’s 6 minutes, the group cycles through 15 choruses of the tune’s 16-bar form while
exploring a wide range of textures and approaches to time, as described in Figure 23.3.

The four melodic phrases that de ne the form of Kern’s composition are mirrored by the ensemble as they C23.P62
shape their performance of the tune. In the rst four choruses, they bring the melodic and harmonic
p. 499 structure gradually into focus through the interaction of bass and trumpet, with saxophone and piano
joining the texture to propel the group into the second quarter of the performance. Rainey’s drums mark the
rst real landing point of the recording, as the group coalesces at the start of the fth chorus before
exploring the melody in increasingly strong and direct fragments. The collective texture expands to the
midpoint of the recording, marked by the start of chorus 8. Throughout this third quarter of the
performance, the group most closely resembles a traditional ‘standards’ ensemble, with the drums and bass
providing a rm rhythmic foundation for the orid horn improvisations. In the nal quarter of the
recording, the group gradually shifts focus and returns to the more pointillistic, phrasal, motivic
exploration with which they began. In this way, Obbligato makes use of the 16-bar, 4-phrase structure of
‘Yesterdays’ as an anchor for both their momentary interactions and the large-scale architecture of their
performance. At the same time, their performance reveals connections to a range of historical and
contemporary styles and techniques.
Figure 23.3

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
Collectively shaping time in Obbligatoʼs ʻYesterdaysʼ. C23.F3

Rainey and his collaborators in Obbligato create original music while articulating a shared connection to C23.P63
p. 500 multiple musical predecessors. The contours of this particular ensemble are di erent from those of
Rainey’s other groups because the musicians choose to apply and coordinate their shared preferences and
in uences in unique ways. Placing themselves within personal stylistic and processual lineages, these
improvisers begin to de ne the musical borders within which they choose to work. Additionally, they
outline the stylistic contours of their performance networks. With a consistent approach to process, stylistic
choices become a primary factor in di erentiating groups from one another and allowing musicians to carve
out individual identities.

As demonstrated in the work of Obbligato, Rainey’s embodiment of the jazz tradition extends far beyond C23.P64
notions of idiomatic style and gesture. Throughout his discussions of his most direct in uences, he shifts
our focus from the stylistic content of musicians’ utterances to the decision-making processes generating
those sounds. In doing so, he frames his own work as an interrogation of the expressive possibilities to be
exploited in each new musical situation. Speaking about some of his favorite jazz drummers, he says:

I think with any of my drummers that mean a lot to me—Elvin [Jones], Tony [Williams], Roy C23.P65
Haynes, on and on—to me, they were the perfect ones for that music. They were that music. They
made that music. My biggest attraction to Tony was just the fact that he really made this new thing
happen. To my ears, it was really fresh and really new. You could hear the lineage; basically, he was
just playing jazz time the whole time, but the way he advanced the rhythmic possibilities of that
music is really fascinating. So it really just comes down to the rhythm.

In emulating the ideals of these older musicians, he strives to be the perfect drummer for his community’s C23.P66
music. In a performance network embracing stylistic eclecticism and improvisational uidity, Rainey
becomes the sound of the community by embodying these traits and tuning his improvisational logic to the
speci c needs of each ensemble and composition.
The Open Loose trio, another of Rainey’s regular working ensembles, provides a di erent avenue for C23.P67
exploration and a di erent improvisational focus than the large-scale compositions of 7 Black Butter ies,
the freeform give-and-take of Paraphrase, or the reimagined standards of Obbligato. Open Loose is a
vehicle for bassist Mark Helias’s compositions, which are generally through-composed pieces between 4
and 8 minutes in length (Helias 2002). Though the group covers a fairly broad range of musical ground, each
individual tune explores a single stylistic area in some depth. Rainey characterizes the group’s focus as
blurring the lines between composed and improvised elements: ‘I’m trying to sustain the mood of the piece,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
the tune itself, to where—this is common in a lot of this music where you don’t want it be necessarily, “OK,
the head’s over and then boom.” Especially Mark’s heads really lend themselves to that because they’re
usually long lines and if you don’t know the tune you’re not really sure if it’s written or not.’

On ‘The white line’, a composition recorded by Open Loose for the 2002 album Verbs of will, Rainey begins C23.P68
by playing a swing pattern with brushes, reminiscent of the straight ballad playing of any number of
mainstream drummers from the late 1950s or early 1960s. At the same time, he updates this standard by
p. 501 uidly interacting with Helias’s rhythmic subdivisions on bass. At the beginning of Malaby’s improvised
saxophone solo, Rainey and Helias work together to create continuity between the written and unwritten
segments of the tune. Discussing his approach to accompanying Malaby’s solo improvisation, Rainey says:

The rst thing is trying to keep that particular feeling of swing to that tune going, but then just C23.P69
kind of adding on to it. Like, there were some little double-time things that we would go into that
never really went into double-time. But Tony might play something that suggested double-time
and I might acknowledge it even though I may not really play exactly what he’s playing, but I’ll
react to it.

As Helias simpli es his bass line, Rainey focuses on playing a simple and consistent pattern on the snare C23.P70
drum with brushes. Rainey quickly begins adding o beat accents, increasingly interrupting the steady ow
of Helias’s walking line through his dissonant placement. As Malaby’s solo progresses, both Helias and
Rainey gradually move away from the dominant pulse and toward dense statements of uneven phrase
lengths. Because of their e orts to maintain a connection between composed and improvised sections, this
eventual dissolution of the pulse never registers as abrupt or jarring.

Throughout its recorded output, Open Loose moves freely between carefully composed and freely C23.P71
improvised modes. Through their shared understanding of process, its members continually blur the lines
between these modes, leaving listeners to wonder which elements are predetermined and which are decided
in the moment. They draw improvisational ideas from Helias’s compositions, amplifying and adapting
those elements in communication. In short, they borrow very well-worn devices from bebop tradition and
practice and apply them in a non-tonal, non-cyclic context. As musicians working broadly within the
context of modern jazz, Rainey and his collaborators draw on particular approaches to gesture, repertoire,
composition, and form as they assert particular connections to traditional practices. Through the speci cs
of their musical histories and shared experience, they thoughtfully expand those traditional practices and
position themselves in relation to jazz communities past and present.
Interacting across time: Conclusion C23.S4

In embodying a musical system based on contextual awareness and strategic deployment of a wide range of C23.P72
stylistic markers and interactive approaches, Rainey and his collaborators mark ‘variety’ and ‘freshness’ as
top priorities, both for themselves and as a responsibility to the musicians with whom they regularly
perform. Over time, di erent ensembles develop unique interactive philosophies, and by playing a variety of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
ensembles, musicians stretch out and develop di erent improvisational ideas. For Rainey, the makeup of a
particular ensemble signi cantly a ects the way he works in the group:

p. 502 I think the biggest distinction [between the music of his working groups…] is actually just the C23.P73
chemistry of the band, and how the band relates. That’s really di erent because the way Tony
[Malaby] relates to composition and improvisation is completely di erent than the way Tim
[Berne] does or Ralph [Alessi] does in Drew [Gress]’s band. So I think the one thing they all share is
that there’s a lot of thought put into the chemistry of the group and then how that chemistry is
going to make whatever sense of the composition.

Together, players tune compositional philosophies, ensemble ethics, and performative patterns to the C23.P74
dynamics of each group as they collectively generate musical time at the scales of gesture and phrase, song
and performance, and lineage and tradition.

Each of his ensembles o ers Rainey the opportunity to explore di erent areas of his improvisational C23.P75
imagination. The personnel of each band and the limitations imposed by composers colour his approaches
to each group, and improvisational strategies internalized in one situation inevitably in uence his work in
other settings. In fact, the certainty of cross-pollination drives the community and keeps its music alive and
vibrant.

Through their shared histories and aesthetic a nities, Rainey and his bandmates maintain strong audible C23.P76
connections to a number of musical communities associated with a variety of genres. At the same time, their
focus on collective improvisational processes allows the musicians in Rainey’s performance network to
connect themselves to a long and uid tradition of play, adaptation, and change. As Rainey and musicians
like him continue stretching, playing, and rede ning the musical structures and timescales within which
they work, they continually nd new resonances in each gesture and across generations.

Notes
1. All interview material with Tom Rainey is drawn from a personal interview conducted on 15 Nov. 2006, unless otherwise C23.N1
stated.

2. All interview material with Drew Gress is drawn from a personal interview conducted on 30 Jan. 2007, unless otherwise C23.N2
stated.

3. See also Johnson (1987). C23.N3

4. See Csíkszentmihályi (1990). C23.N4


References C23.S5

Berne, T. (2002). Science friction. CD 013. Screwgun. C23.P77


Google Scholar Google Preview WorldCat COPAC

Berne, T. (2004). Electric and acoustic hard cell live. CD SC70014. Screwgun. C23.P78
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471201 by National Science & Technology Library user on 26 May 2023
Borgo, D. (2006). Sync or swarm: Improvising music in a complex age. Continuum. C23.P79
Google Scholar Google Preview WorldCat COPAC

Bowman, W. (2004). Cognition and the body: Perspectives from music education. In L. Bresler (ed.), Knowing bodies, moving C23.P80
minds: Towards embodied teaching and learning, 29–50. Kluwer Academic.
Google Scholar Google Preview WorldCat COPAC

p. 503 Chinen, N. (n.d). Tom Rainey talks with Nate Chinen. Screwgun. C23.P81
https://web.archive.org/web/20120805223118/http://www.screwgunrecords.com/page_a.php?pageid=interviews&sub=rainey
Google Scholar Google Preview WorldCat COPAC

Csíkszentmihályi, M. (1990). Flow: The psychology of optimal experience. Harper & Row. C23.P82
Google Scholar Google Preview WorldCat COPAC

Gress, D. (2005). 7 Black Butterflies. CD 5809. Koch/Premonition. C23.P83


Google Scholar Google Preview WorldCat COPAC

Helias, M. (2002). Verbs of will. Open Loose. CD 11. Radio Legs. C23.P84
Google Scholar Google Preview WorldCat COPAC

Johnson, M. (1987). The body in the mind: The bodily basis of meaning, imagination, and reason. University of Chicago Press. C23.P85
Google Scholar Google Preview WorldCat COPAC

Laskey, K. (2017). On the standards playground: Tom Rainey speaks. Jazzspeaks.org. http://www.jazzspeaks.org/on-the- C23.P86
standards-playground-tom-rainey-speaks/
WorldCat

Paraphrase (1997). Visitation rites. CD 70002 Screwgun. C23.P87


Google Scholar Google Preview WorldCat COPAC

Rainey, T. (2014). Obbligato. CD 227. Intakt. C23.P88


Google Scholar Google Preview WorldCat COPAC

Urban, G. (2001). Metaculture: How culture moves through the world. University of Minnesota Press. C23.P89
Google Scholar Google Preview WorldCat COPAC

Varela, F. J., Thompson, E., and Rosch, E. (1991). The embodied mind: Cognitive science and human experience. MIT Press. C23.P90
Google Scholar Google Preview WorldCat COPAC

p. 504 Whitehead, K. (2014). Tom Rainey Obbligato. Liner notes. CD 227. Intakt Records. C23.P91
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
CHAPTER

24 Time and Ensemble Dynamics in Indeterminacy:


C24 John
Cage’s Concert for Piano and Orchestra 
Emily Payne

https://doi.org/10.1093/oxfordhb/9780190947279.013.17 Pages 505–C24.P109


Published: 08 December 2021

Abstract
This chapter examines ensemble dynamics and time consciousness in indeterminate music, using John
Cage’s Concert for Piano and Orchestra (1957–8) as a case study. Drawing on interviews and
observational studies undertaken with the experimental music ensemble Apartment House, I examine
the role of temporal indeterminacy in the socio-musical interactions that characterize performance,
and its implications for the musicians’ experiences. In doing so, the chapter makes a broader
contribution in its consideration of the ways in which issues of authorship and authority are negotiated
in such temporal interactions, and how the dynamics of these negotiations present a sociality based on
a ‘separate togetherness’, whereby performers play together (out of time) with one another.

Keywords: John Cage, conductor, indeterminacy, time, temporality, ensemble dynamics, sociality
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Introduction C24.S1

PERFORMING together in an ensemble is typically associated with principles of communication and C24.P1
intersubjectivity—a coming together in a shared experience as musical events unfold over time. Of all
ensemble formats, the orchestra in particular has frequently been used as a metaphor for sociality, in which
musicians work together to create a coordinated performance. Orchestral practice has been described by

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
Robert Faulkner (1973) as representing ‘an exemplary process of collective action-making’ (p. 147); writing
more recently, Melissa Dobson and Helena Gaunt (2015) characterize orchestral musicians’ actions as being
‘formed as a result of adapting and synchronizing to any number of their colleagues’ (p. 31). While the
individual and collective experiences of orchestral musicians are clearly made up of a complex constellation
of behaviours, decisions, feelings, and so on (Cottrell 2017; Ponchione-Bailey 2016), there is a general
consensus that most orchestral musicians attend to the ‘prime directive’ (Ponchione-Bailey 2016) of
playing together in a coordinated and cohesive manner, guided by a uni ed group intention. But what
happens when that unity does not exist, whether through chance, or deliberate resistance?

This chapter explores how time and temporality are experienced in the performance of indeterminate C24.P2
music, using John Cage’s Concert for Piano and Orchestra (1957–8) as a case study. Cage’s interest in time has
p. 506 been well documented (Campana 2001; Pritchett 1993; Valkenburg 2010), and his concern to explore the
organizational structure of the orchestra is evident in a number of other works throughout his
1
compositional output. Each performance of the Concert exhibits multiple temporal strata that recon gure
musical time on several levels: by introducing indeterminacies to musical structure, at the level of the page,
system, and event; by removing conventional systems of metric time; and by providing a performance part
for conductor, who functions as, in Cage’s words, ‘a chronometer of variable speed’ (Cage et al.
1992b[1958]) by altering the pace of the performance in unpredictable ways. Drawing on interviews and
observational studies undertaken with the experimental music ensemble Apartment House, I examine the
role of temporal indeterminacy in the socio-musical interactions that characterize performance, and its
implications for the musicians’ experiences. In doing so, the chapter makes a broader contribution in its
consideration of the ways in which issues of authorship and authority are negotiated in such interactions,
and how the dynamics of such negotiations present a sociality based on a ‘separate togetherness’ (Iddon
2
and Thomas 2020: 190), whereby performers are invited to play (out of time) with one another.
Ensemble Dynamics and Time Consciousness in Musical Performance C24.S2

A key question—perhaps a deceptively simple one—that underpins this chapter is: ‘What do musicians C24.P3
experience during performance?’ Discussions of attention and awareness in performance often turn to the
phenomenon of ‘ ow’ (Csíkszentmihályi 1996)—a term used to describe a heightened state of
consciousness that occurs during peak experiences whereby the practitioner is fully absorbed in the task at

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
hand. This apparent e ortless automation is not xed, however. It is borne out of a dynamic interplay of
body, instrument, and skill, the outcome of an optimal relationship between challenge and capacity that
enables performers to anticipate and deal with ambiguity and resistance in the moment. In his research on
group creativity, Keith Sawyer (2003) argues that while ‘intersubjectivity is fundamentally social and
collective’ it must be ‘negotiated’ (p. 24). Invoking Mihály Csíkszentmihályi’s work, Sawyer proposes the
phenomenon of ‘group ow’—that is, the shared attentional awareness of a group of musicians that can
occur when they become collectively attuned, whereby ‘everything seems to come naturally’ and ‘the
performers are in interactional synchrony’ (Sawyer 2003: 44). Group ow, Sawyer argues, occurs when a
group achieves a balance between the unpredictability of their improvised actions and the coherence of their
shared knowledge and skills. Although participants might experience their own ow states simultaneously,
this is not a prerequisite for group ow: it is an emergent and irreducible phenomenon of a performing
ensemble that can only occur when equal contributions are made by each participant (see also Sawyer 2014:
273).

p. 507 While Sawyer is primarily concerned with jazz improvisation, Frederick Seddon and colleagues (Seddon C24.P4
2004; Seddon and Biasutti 2009; see also Waddington 2017) identify a similar mode of communication in
both jazz and classical performance, de ned as ‘empathetic attunement’: the phenomenon of musicians
engaging with one another during performance. Empathetic attunement holds that musicians must ‘de-
centre’ (that is, shift their focus away from themselves and onto their co-performers) in order to have
shared common goals and to ‘see things from other musicians’ musical points of view’ (Seddon and
Biasutti, 2009: 120). This fundamental intersubjectivity of music-making lies at the heart of Nicholas
Cook’s conception of music as performance (Cook 2007[2004]). Cook draws on the sociologist Alfred
Schütz’s (1951) concept of ‘inner time’ to argue that all performance is a matter of social interaction, a
‘mutual tuning-in’ (Schütz 1951: 79) of performers that leads to ‘a shared, communal temporality’ (Cook
3
2007 [2004]:15). That is to say, all musical performance a ords a collective communality; an argument
that Cook uses to critique ontological understandings of composition, performance, and improvisation.

The above characterizations of ensemble relations view the experiences of performers through a lens C24.P5
focused primarily on principles of convergent coordination, an approach that is predicated on collective
attunement. In his phenomenological account of absorption in musical performance, Simon Hø ding
(2019) has questioned the idea that players need to attend to each other’s states of consciousness during
performance. Through detailed ethnographic work with the Danish String Quartet, Hø ding shows how a
performance might be achieved with players occupied by, or oblivious to, all manner of di erent
considerations (both musical and non-musical), and not necessarily attuned to how their colleagues are
feeling, or what they are experiencing or attending to. Instead, he suggests that, while performers might
have rehearsed and learned a particular interpretive approach, how that performance is realized through
playing might uctuate (even if imperceptibly) in unpredictable ways, with each musician’s sense of agency
in the unfolding music shifting in various ways. For example, he categorizes performers’ experiences as
oscillating between moments of ‘standard absorption’ (when a player’s expectation and execution are
perceived to be equitable), mind-wandering, ‘frustrated playing’ (on encountering a particular obstacle),
‘intense absorption’ resulting in a perceived loss of self, and ‘ex-static’ absorption (a form of re ective
consciousness accompanied by a heightened state of agency) (Hø ding 2019: ch. 4). Hø ding’s view of
ensemble performance is therefore emergent and interactive, but not predicated on uni ed or collective
experience. This argument thus accounts for a more di erentiated and dynamic view of co-performer
interaction—one which does not rely on shared subjectivity, whether that interaction takes place between
members of a string quartet or a symphony orchestra.

Rather than occupying a continuously shared ‘inner time’ then (pace Schütz), there might be several C24.P6
entangled streams of sensations and perceptions at play in a performance as players negotiate and move
between di erent kinds of interactions. The sociality of performance is therefore not unmediated or neutral,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
but is culturally shaped and experienced. Mark Do man (2019) makes precisely this argument in a
p. 508 discussion of the ways in which musicians inhabit time in the moment of performance—i.e. their
attention to, and awareness of, the when in performance. Do man describes time consciousness as an
‘outward-facing, […] socially-oriented practice that takes account of others’ actions and which musicians
both inhabit and act upon’, in a way that incorporates ‘the complex of awarenesses and their
transformation within music’ (p. 184). Using the Greek division of time into chronos (‘processual’) and
kairos (‘event’), Do man argues that performance requires a combination of both kinds of temporal
awareness. Akin to Schütz’s ‘inner time’, processual time consciousness describes a musician’s tacit and
encultured awareness of the temporal character (i.e. the metre, rhythm, and tempo) of a particular style or
genre of music, and thus what it means to play in time in that music. Event consciousness is the awareness
of timely or opportune moments during a performance, the sense of when something should occur (in a
rhetorical sense) that often occurs at pivotal moments in performance (Do man cites the ending of a jazz
improvisation or the punchline of a joke as particularly apt examples of event consciousness). However, in
contrast to Schütz’s model, processual time is not uniform or unmediated. Time consciousness, Do man
argues, operates at di erent, but interrelated, scales of engagement which are located both ‘inside’ and
4
‘outside’ the room. At the narrowest, micro level, a performer might be immediately aware of what they are
doing in the moment of a performance (whether in relation to their own playing and/or in relation to their
colleagues); at a broader, meso level, a performer’s time consciousness is shaped by their enculturation
within a tradition or community of practice (Wenger 1998); and at the broadest, macro level, a musician’s
time consciousness must adapt to paradigm-shifting changes within a domain (e.g. innovations such as the
invention of the metronome or click track). During performance, players’ momentary awareness of time
(or, as Do man puts it, their practical time consciousness), is shaped by particular discursive practices
which operate from the micro to the macro level.

If the preceding discussion seems to suggest that spontaneity and contingency are inherent to all ensembles C24.P7
(even in the most ostensibly ‘mechanistic’ musical situations such as musical theatre pit bands), rather
than being a unique aspect of indeterminate music, on one level this is true: Hø ding’s characterization of
absorption certainly resists a view of performance as being any more or less determined by a score. But it is
important also to take into account the particular performance practices associated with Cage’s music, and
the Concert speci cally (and thus attend to the meso and macro level in uences). The following section sets
out some of those practices in more detail.
ʻ…a Solo or a Part in an Ensemble, Symphony, or Concertoʼ C24.S3

5
The somewhat ambiguous ‘concert’ of the piece’s title, as well as its instructions for each instrumental part C24.P8
p. 509 to be conceived of as ‘a solo or a part in an ensemble, symphony, or concerto’ (1960 [1958]), hint at
Cage’s reframing of a conventional ensemble situation. This is also suggested by the performance materials:
the piano part, also known as the Solo for Piano, comprises 84 notations spread across 63 large unbound

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
pages, which can be played in any combination and any order. The orchestral musicians play from separate
parts (without a score), and, working within a predetermined length of time, each performer can choose to
play any number of pages of their part (including none). Moreover, the Concert’s formal instructions state
that the parts themselves can be played in any combination, including with other pieces, o ering seemingly
6
endless performance possibilities.

The notations and instructions of each part follow more or less similar principles, particularly with regard C24.P9
7
to timing. The length of each system (or stave) is unspeci ed, and all sounds are indicated using either a
small, medium, or large note head. Cage’s instructions state that these three sizes determine either the
amplitude or the duration of a particular pitch (i.e. a small note head can be of either a quiet dynamic, short
duration, or both, and so on). While no direction is provided for how to decide the placement of notes across
a system, in the programme notes for the Concert’s premiere Cage speci ed that ‘each part is written in
detail in a notation where the space is relative to time determined by the performer and later altered by a
conductor’ (Cage and Kostelanetz 2000[1993]: 57). The player must therefore negotiate the spatio-
temporality of each system relative to the potential duration of each sound.

Performances of the Concert do not necessarily require a conductor—the instructions suggest that the C24.P10
musicians could use a stopwatch or similar device (although this would result in quite a di erent, less
exible performance); if a conductor is involved, however, they work from their own part too, which
8
provides a list of timings to (in Cage’s words) change ‘clock time to e ective time’ (see Figure 24.1).

During performance the conductor uses both arms to imitate the hands of a clock. What this means in C24.P11
practice is that, standing in front of the orchestra in the usual position, the conductor begins with their left
arm stretched up above the top of their head at the 12 o’clock position. Their arm moves around in an anti-
clockwise (from the conductor’s perspective) motion like the second hand of a clock; when it reaches the
bottom (6 o’clock) point, their right arm continues on the same trajectory to return to the top (12 o’clock)
position, completing one full rotation for one minute of ‘e ective’ time. The pace of these movements
uctuates according to the conversions speci ed in the conductor’s part, anywhere from eight times slower
to two times faster than normal ‘clock’ time. The orchestra musicians must follow the timings set by the
conductor, moving faster or slower accordingly, with conventional performance practice holding that the
pianist ignores the conductor’s actions. In addition, the conductor does not necessarily need to start from
the rst row of the part, and so the actual sequence of timings might be quite di erent from one
performance to another. In the Apartment House performance discussed in this chapter, the musicians
prepared 39 minutes of material to correspond to 39 revolutions of the conductor’s arms, which—with the
9
p. 510 uctuations in speed of the revolutions in the actual performance—lasted just over 53 minutes.
Figure 24.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
John Cage: Concert for Piano and Orchestra (1957–8), Conductorʼs part, p. 1. Edition Peters No. 6705. © 1957 by Henmar Press C24.F1
Inc., New York.

Reproduced by permission of Peters Edition Limited, London.

p. 511 Cage described the role of the conductor in the Concert as being ‘Not of a conductor or a leader, but of a C24.P12
utility!’ (Cage and Charles 1981[1976]: 109). On the one hand, the Concert reduces the conductor’s role to a
single function: controlling time. Yet, on the other, their authority becomes more absolute, since the
relationship between conductor and ensemble is not reciprocal. The performers must respond to the
conductor’s actions, but the conductor does not visibly react to the musicians or indeed to any aspect of the
music as it unfolds. In one sense, then, the conductor ‘directs’ the ensemble in a one-directional manner,
with the pianist working alongside, yet independent from the events on stage. At the same time, the
conductor exercises a sort of authority ‘by proxy’ since the authority is not their own: they must adhere to
the strictures of the conductor’s part.

This approach to indeterminacy in performance is summed up by Cage as the ‘practicality of anarchy’ (1972, C24.P13
cited in Kostelanetz 2003[1987]: 67), and is operationalized in the Concert on a number of levels. The rst is
by disrupting the performers’ embodied relationships to their instruments. Instruments are dismantled,
detuned, and destabilized, and within the parts themselves techniques are often stretched or combined to
the point of complete breakdown. Second, and perhaps most profoundly, indeterminacy unsettles the
temporal relationships between the ensemble members. Each instrumentalist has a sense of what is going
to happen in terms of the order of events on their individual part, but they have been denied information
about when it is going to happen, as well as what is going on with the rest of the ensemble, i.e. the unfolding
sound world around them. They do not need to maintain an agreed tempo or attend to metre, but they do
need to watch the conductor and be sensitive to uctuations in e ective time in order to gauge the
placement of events, and to adjust the durations of sounds accordingly. Conversely, the conductor can

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
anticipate the overall duration of a performance and any uctuations in speed, but the what and the when of
musical events is completely unknown. What sets this temporal landscape apart from the stretching and
placement required during a more ‘conventional’ concert hall situation is that there is no mutual
relationship between the players, through either the existence of a shared pulse or a collective sense of
architecture.

In considering how a performance of the Concert might di er from a performance of a piece of Western C24.P14
(classical) music, G. Douglas Barrett’s (2016) description of the score as ‘less a blueprint that mandates a
preconstructed musical object and more a prompt that produces a series of contingent consequences in its
realization’ (pp. 48–49) is apt here. In the case of the Concert, the performer’s intimacy with their material
is undermined by the contingencies of the performance situation which deny them certain information. The
performer can work closely with the material of their part so that the various events and actions (including
often quite complex techniques) become thoroughly ingrained within their practice. However, while the
performer might have a comprehensive understanding of the timings of their individual part, during the
performance itself their expectations might be frustrated by the unpredictable and unstable shifts in speed
brought about by the conductor’s actions. Iddon and Thomas (2020) have characterized these conditions as
resulting in a defamiliarization of the performer’s habitual actions, whereby:

p. 512 the conducted changes of pace demand shifts in energy or movement [which] might be more-or- C24.P15
less nebulously felt, but wholly tangible would be the di erent ways in which some actions must be
articulated at an unexpectedly faster or slower pace or, equally, the way in which, should a sudden,
unexpected accelerando take place the way in which a particular, well-known musical event might
need to be dropped entirely or, in the case of an equally unexpected ritardando, an event perhaps
not seemingly amenable to extension has to be somehow stretched or otherwise drawn out. (p. 98)

That is to say, the temporal indeterminacy of the Concert has perceptual consequences for each musician’s C24.P16
‘object of expectation’ (Huron 2006: 6), and thus disrupts their potential immersion in the ow of
performance in more or less discernible ways. While it would be problematic to take an overly deterministic
view of these circumstances, such that the performers are entirely constrained by Cage’s notation, it could
be argued that the performance situation places selective pressures on the musicians that must be dealt
with through playing. Another way of characterizing this situation might be to view it as an instance of
‘improvisation impromptu’ (Goehr 2016), which describes the phenomenon of grappling with unexpected
10
resistances in the moment of performance, since the musicians must adapt their playing ‘on the y’ to
abrupt changes in the music, including adjustments to speed such as those discussed above. In this way, the
performers’ embodied knowledge of the music must be su ciently malleable to adjust to unanticipated
constraints in the moment of performance.

Clearly, the phenomenon of musicians having to ‘think on their feet’ through playing is not the preserve of C24.P17
indeterminate music. Accounts from performers show that a sense of risk and spontaneity can be
experienced in even the most apparently ‘determinate’ performance situations, such that ‘just about
anything can happen’, to quote violist and member of the Guarneri String Quartet, Michael Tree (cited in
Blum 1986: 20). Moreover, ensembles routinely adapt their playing to maintain and restore synchrony
during performance (Weeks 1990; 1996). The di erence lies in the way that, in the context of the Concert,
disruption and ux are, if not actively pursued, then at least accepted aesthetic ideals, rather than breaks in
continuity that must be resolved. A corollary of these circumstances is that, as William Brooks (2002) has
argued, the performer’s ‘responsibility is to themselves and to their own actions, not to an imagined
totality’ (p. 222). It is also possible, of course, that the conductor could share with the other musicians the
changes of speed in advance of a performance of the Concert, and that the performance itself could be
thoroughly rehearsed to the point of relative predictability. It is arguable that such an approach would have
deviated so far from indeterminacy’s aesthetic of celebrating di erence and contingency as to be contrary

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
to the spirit of the performance tradition. Moreover, aside from the fact that it would be quite an impractical
method to execute, given that Cage could have speci ed instructions for such a practice, it seems telling that
he chose not to. Nevertheless, this is precisely the approach taken by the pianist and rst performer of the
Concert, David Tudor, whose meticulous yet sometimes idiosyncratic realizations of Cage’s scores
p. 513 e ectively ossi ed them into ‘determinate’ pieces, and yet whose performance practice Cage celebrated
(Iddon 2013; Pritchett 2004). This paradox illustrates one of the inherent tensions between theory and
practice in understandings of indeterminacy (Cline 2019), while also emphasizing that the case study
discussed in this chapter presents just one of several possible interpretations of Cage’s music.

Performing the ʻPracticality of Anarchyʼ C24.S4

A conventional view of a Cageian performance practice associated with his music from the 1950s onwards is C24.P18
one characterized by individualism and isolation, requiring performers to work independently from one
another, and to have the self-discipline to refrain from making decisions based on personal choice. Such an
approach is suggested by Cage in a letter concerning Variations III (1962–3), where he advocated a
‘disciplined’ mode of performance, i.e. ‘a way not constrained by [the musician’s] likes and dislikes’, and
that rehearsals should ‘be in solitude, not together with any of the other musicians’ so that ‘[n]o one is to
11
constrain or be constrained by another’ (Cage, personal communication to Charles Boone, 21 March 1968).
A similar view is articulated in Cage’s recollection of his response to an invitation to play with the Joseph
Jarman Quartet in 1965, in which he ‘advised them not to listen to each other, and asked each one to play as
a soloist, as if he were the only one in the world’, appealing to the musicians that they should not respond to
one another’s dynamic levels; rather, ‘they should be independent, no matter what happened’ (Cage and
Charles 1981[1976]: 171; see also Kim 2012). Cage’s suggestions here are focused on resisting
intersubjectivity and the tuning-in commonly ascribed to ensemble performance.

One method of facilitating autonomous performer behaviour, according to Cage, was through the physical C24.P19
separation of the musicians in the performance space. Cage spoke about this in some detail in his lecture
‘Indeterminacy’, delivered in Darmstadt in the same year as he nished composing the Concert. He argued
that the spatialization of seating in this way would allow for ‘the independent action of each performer’
since ‘[t]here is the possibility when people are crowded together that they will act like sheep rather than
nobly’ (Cage 1961b[c.1959]: 39). Cage’s statement here is concerned with the ‘chronic’, or processual time
consciousness, the coordinative impulse to play together, an impulse that his comments suggest should be
repressed. Separating the musicians from one another brings the performer’s taken-for-granted, ock-like
awareness into focus, and a ords an ‘interpenetration’ rather than a ‘harmonious fusion’ (p. 40) of sounds.
Writing from a performer’s perspective, Petr Kotík, the conductor and director of the S.E.M. Ensemble, who
has performed Cage’s music widely, places a similar emphasis on self-discipline—understood here, like
Cage, to mean an aesthetic that emphasizes individual responsibility—in this example in relation to the
p. 514 Concert itself. Performance choices, Kotík argues, ‘must not be made merely according to the performer’s
preferences’ so as to ‘remove value judgments from the decision-making process’. He sums up decisively,
‘One should never put one’s own ego between the music and the performance!’ (Kotík 1993).
In order to distance their choices from relying on habit or musical training, musicians might employ chance C24.P20
procedures as part of their preparation process, such as tossing a coin, using a random number generator,
deriving operations from the I Ching, and so on. But in terms of how musicians respond to one another in
the moment of performance, the question remains about the extent to which it is practically possible to
‘switch o ’ one’s awareness from one’s co-performers. The situation of performing the Concert is certainly
less clear-cut than Cage’s or Kotík’s comments suggest, with its contingencies resulting in (from the
performer’s perspective) certain thresholds being shifted from those of conventional ‘concert’ music. This

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
relates to its emphasis on performers as individuals, through the use of separate performance parts (and the
eschewing of a ‘master’ score), and through the physical dispersal of the players within the performance
12
space. On the one hand, the ensemble could be understood as comprising a collective of soloists rather
than a uni ed group, with each player operating independently of one another: there is no need for any
player to respond to, or communicate with, their colleagues. On the other hand, the presence of a conductor
invites a ‘mutual tuning in’, both symbolically, in creating a more conventional, almost parodic, version of a
concert hall situation, and in practical terms: the conductor provides the musicians with a shared point of
reference both visually and temporally. As Iddon and Thomas (2020) put it:

[T]he conductor reintroduces—if transformed—one of the usual strictures of large ensemble C24.P21
music: with a conductor, the ensemble is no longer simply a group of co-existing soloists, but an
assemblage of intersecting individuals who exist in collective temporal space. (p. 206, emphasis
original)

These conditions lead to performers inhabiting a ‘separate togetherness’ (190), whereby at a group level C24.P22
performers must operate quasi-collectively to respond to the changes in speed set by the conductor, while
at an individual level they operate independently.

A performance of the Concert thus treads a line between presenting a collection of instrumentalists who are C24.P23
apparently hermetically sealed o from one another and mimicking a typical orchestral situation. But how
does this tension play out in practice? The conduct of the musicians during the rst performance of the
Concert at New York Town Hall on 15 May 1958 forms part of a colourful narrative surrounding the piece, and
does not need to be repeated here; su ce to say that early performances were characterized by the allegedly
disruptive behaviour of the orchestral musicians, and similarly playful—or, in some cases, rowdy—
13
responses from the audience. Despite this infamous reception history, however, there are few rst-hand
accounts from the musicians themselves about what it was like to perform the Concert. The extent to which a
p. 515 typical Cageian performance ideology relates to the actual experiences of musicians working within the
practical constraints of performance is the focus of the subsequent case study.

Case Study: Apartment House C24.S5

The case study addressed in this chapter is based on a performance by Apartment House conducted by Jack C24.P24
Sheen with Philip Thomas playing the piano part, which took place on 1 July 2017 at the University of Leeds
and resulted in a CD recording (Apartment House 2017). A rehearsal on the day of the performance and the
14
performance itself were lmed. Semi-structured interviews were undertaken with the musicians
15
throughout the project, both before and after the performance.

At the macro-social level, the Concert o ers two paradigm shifts in orchestral performance practice. The C24.P25
rst is its rejection of a metric temporality which would usually facilitate rhythmic perception and action.
The music requires the elimination of the metric measuring of time, where musical events unfold in seconds
and minutes rather than beats and bars. In practice, this means that players do not need to tune in to a
shared pulse. Indeed, the main challenge that Sheen identi ed during performance was being able to orient
himself within the music without relying on listening to the ensemble’s sounds, commenting during
interview:

You’re totally divorced from the players’ music, however they are in some way connected to you. C24.P26
[…] Like all musicians, we’re trained to tune in to sounds, and this was a piece where if I did that I
knew I’d get lost in my own counting and what I was doing, which bears no relationship to the
16
sounds in a way that I need to be aware of.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
Sheen’s shift in focus had an impact on his recollection of certain details of the performance. He described C24.P27
the experience of performing as being ‘like a very broken up piece of music, but each section of that requires
a lot of attention. It kind of erases your overall sense of architecture, and actually the piece ies by because
17
of that.’ For cellist Anton Lukoszevieze, the experience of playing with others in a state of metric
dislocation disrupted his ability to concentrate on his own part, and he strove to avoid listening to the other
players so as not to have his attention diverted from his playing. As with Sheen, this had an e ect on his
perception of the passage of time, and he described the experience as ‘quite disorientating […] it’s actually
quite exhausting just sitting there waiting to play, because you don’t want to be distracted and you really
18
just have to prepare yourself in your mind about what you have to do that’s coming up.’ As a consequence,
some players decided to predetermine the details of their parts as part of their preparation process, leaving
little to ‘chance’ in performance. Violist Reiad Chibah, for example, asserted that it was important to
p. 516 prepare a relatively xed realization, in order to suppress his ‘instinct to respond to sounds going on
19
around’ him, and thus to facilitate a sense of separation from his colleagues.

The second paradigm shift manifest in the Concert is its recon guration of the role of the conductor, and the C24.P28
consequences of this for (1) the musical interactions of the musicians (since the passing of the seconds and
minutes is disrupted by the conductor’s actions) and (2) their social interactions in the unsettling of the
conductor–orchestra relationship. At the meso level, each performer’s sense of timekeeping and timeliness
was in uenced by their personal experiences, training, and knowledge of the experimental music tradition
that form part of Apartment House’s collective aesthetic and conduct, as well as the ensemble’s experiences
of rehearsing and performing together. For example, other than discussing their own personal performance
strategies with the research team during interviews, the group chose not to consider potential preparation
or performance strategies collectively. Moreover, as I will show, while all of the musicians demonstrated
careful attention to performing their parts, they each placed a di erent emphasis on Cage’s instructions.
Some brought an understanding of Cageian performance practice to bear on what they did, but there were
inevitable di erences in what a ‘Cageian’ performance practice was understood to mean, how that might sit
alongside other performance priorities, and, perhaps somewhat inevitably, how those priorities played out
in the moment of performance.

Within the ensemble there was a tendency for the performers to experience surprise at the di erent sounds C24.P29
that they encountered as they heard the Concert take shape as a collective performance for the rst time. In
direct contravention of Cage’s comments, some performers described feeling drawn to respond to the
sounds of their colleagues and thus adapting their choices in the moment of performance. This might mean
extending a quiet sound for longer so that it could be heard in the midst of particularly noisy activity, or
holding back from a loud and intrusive entry during an otherwise quiet episode. Other players enjoyed the
feeling of their playing suddenly being more prominent in the collective sound world. For Chibah, for
example, this experience was heightened by the layout of the performance space, with the separation of
20
musicians creating a ‘transparency’ in the sound world that let each player present ‘points of interest’
within moments of silence in a way that he had not experienced in other performances.
A ʻNon-musicianʼ C24.S6

The Concert’s shift in time consciousness at a broad, macro-social level is exempli ed in Sheen’s re ections C24.P30
21
on the unconventional nature of his role, one which he described in interview as ‘very “unmusical” ’.
Indeed, from the conductor’s point of view, there is no ‘music’ as such to learn: the conductor’s part only
indicates how e ective time and clock time relate to one another. A telling point is that Sheen did not listen
to a recording of the Concert or look at any of the instrumental parts before the performance. The

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
ensemble’s brief rehearsal was used to perform a sound check and to make sure that all members of the
p. 517 ensemble understood how the rotations worked, and how the piece would begin and end, but even at this
point the piece was not performed in its entirety. Sheen described his function as essentially one of ‘time
22
management’, a description reinforced by violinist Ruth Ehrlich. She contrasted Sheen’s function in the
Concert with the expected function of a conductor: ‘[W]hat he was doing in this performance was certainly
23
not beating time. He was just indicating the passage of blocks of time.’ In addition to recalling Cage’s
assertion that ‘[b]eating time is not necessary. All that is necessary is a slight suggestion of time’ (Cage
1961b[c.1959]: 40), Ehrlich’s comment underlines the expectation that a conductor would usually actively
participate in the orchestra’s shared ‘prime directive’ by communicating the musical beats to the ensemble.

At the meso level, Sheen’s approach to his role was also in uenced by his understanding of the C24.P31
experimental tradition and the performance history of the Concert. In addition to emphasizing the
conductor’s functionality, Sheen saw his role as highly theatrical, in uenced in part by his knowledge that
the dancer and choreographer Merce Cunningham conducted the Concert at its premiere, and that it was
sometimes used to accompany Cunningham’s choreography for the dance Antic Meet (1958). Consequently,
his preparation for the performance was focused on learning what he described as a ‘choreography’—i.e. ‘a
series of physical gestures and adopting a certain physical presence on the stage, which in my opinion,
substitutes for the musicality that a conductor would often give to an ensemble and therefore give to an
24
audience’. Gestural communication is clearly an integral part of all conducting technique, but in this role,
Sheen’s approach to gesture, eye contact, and so on was designed to resist communicating with the
ensemble. He described feeling conscious to not ‘look at players and give them a sense of expectation or
25
imply anything. It’s kind of removing myself from as much of it as possible.’ In this way, Sheen exercised
the role of a quasi-performer, with his attention directed primarily towards the physical aspects of his own
26
individual performance and how they were conveyed to the audience, rather than to the ensemble.

For several players, the absence of the conventional communicative cues and gestures that would be usually C24.P32
expected from a conductor resulted in confusion, demonstrating the consequences of this shift in time
consciousness for the narrower, micro-social aspects of performance. For example, violinist Hilary Sturt
re ected in interview:

[I]t’s the non-involvement, because as soon as somebody stands in the middle of an ensemble, as C24.P33
a ‘conductor’, I know we shouldn’t, but we pay attention to them, and we expect information from
them as to how to behave. But in fact [Sheen] was the reverse. He was a non-information, he was a
non-musician, which is a huge inversion of roles. And normally you’d expect a conductor to
conduct: to give gestures, and movements, and eye-contact, and to change balance. But nothing of
27
that was there. It was a beautiful reversal.

Reinforcing Sturt’s assessment of events, some musicians described becoming disoriented during the C24.P34
performance, and losing track of their place in the performance in relation to Sheen’s rotations. Sturt’s
semi-confessional ‘I know we shouldn’t’ is a tacit acknowledgement of the tension that exists between
p. 518 ideology and practice in this situation. Her comments suggest that a conductor standing before an
orchestra represents a collective attentional focus and creates an intuitive sense of togetherness for the
ensemble; and yet here the Concert’s suppression of the relationship with the conductor had an alienating
e ect for some musicians. This is not to say that confusion and dislocation are not characteristics of any
other—particularly unrehearsed—orchestral situation, but that here those conditions are deliberately
pursued, rather than being symptomatic of a lack of performance preparation or experience.

An Agonistic Sociality C24.S7

Some of the performers’ experiences of performing the Concert were suggestive of the aforementioned C24.P35
agonistic mode of improvisation impromptu put forward by Goehr (2016), in that the changes in pace of the

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
‘conductor minutes’ introduced additional constraints to work within in real time. Clarinettist Vicky Wright
re ected on the disruptive impact that the sometimes abrupt shifts in speed had on her performance,
forcing her to adapt her material in the moment:

Sometimes the minutes were elongated, so what was a minute would have more than doubled, and C24.P36
other times they went quickly, and it really made me re-evaluate my material. I found myself
sometimes missing out things I didn’t have time to get through. And I found myself extending
28
things that previously I had not given that much time for.

Other musicians, such as tubaist Melvyn Poore, echoed Wright’s observations, commenting on having to C24.P37
extend sounds for as long as possible until running out of breath and having to abandon them completely,
and expressing acceptance about the fact that these outcomes were simply out of his control.

Despite these tensions being exercised ‘behind the scenes’ of the performance, neither the conductor nor C24.P38
29
the ensemble responded to one another in any observable way during performance. Sheen described
experiencing a playful sort of ‘being in the know’ in his anticipation of changes in speed and how they
would impact on the players’ actions:

I’m in a privileged position because I know how fast the rotations are about to go, and the players C24.P39
don’t. […] [T]here were some great moments where certain members of the group would start
playing a sound that had a particular character to it, i.e., it was loud or long or it was particularly
unusual, and I knew when they started making that, that most likely without them realising it,
they’d have to sustain that for a very long time, but I was in on it and they weren’t. And that’s so
interesting, and that’s really satisfying, because you’re aware of the impact that your physical
30
gestures are having.

Sheen’s experience here is marked by a sense of friendly antagonism and agency in exerting in uence over C24.P40
p. 519 the musicians’ actions, yet there was also a sense of spontaneity and discovery as he encountered the
players’ musical material in unexpected ways, as if hearing the piece for the rst time. Sheen’s comments
thus shed light on the distinctive function of the conductor in this performance and its impact on his usual
modes of interaction at the micro-social level. They are also indicative of the piece’s recon guring of the
conventional power dynamics between conductor and ensemble.

While violist Bridget Carey also noted that the Concert’s temporal indeterminacy led to moments of feeling C24.P41
unsettled during the performance, she was somewhat cautious about overstating the impact of the Concert’s
temporal instability on her as a performer in relation to other music. In the following comment she
describes experiencing a sense of playful awareness in encountering and adapting to unforeseen obstacles
in the moment of performance:

Composers mess with time in all sorts of ways. This is a radical example of that, but it’s not a C24.P42
counterintuitive thing for a musician to do. And to read the score graphically, over this notion of
changing time is in some ways quite liberating. There were moments when I started to play a note,
and I realized I was going to have to play that note for a very, very long time! And sometimes it’s
not necessarily the most comfortable thing in the world. But also it was quite good fun from that
point of view. ‘Oh right! OK! This is the game now! This is the problem that I need to solve at this
point in the performance.’ And that’s not an unusual thing for a musician to do either. That’s what
musicians do: they solve the problem in relation to the score in the course of a performance. That’s
31
a bit of a job description.

Carey’s comments parallel those of Sheen in describing the e ect of the Concert’s severing of the reciprocal C24.P43
relationship between conductor and ensemble, leading to a grappling or playing with authority in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
performance, as evidenced by their internalized interactions with one another. The dynamic becomes
something of a benign struggle, with neither musician quite sure how (or if) events will resolve. Carey and
Sheen’s temporal experiences show that there is a pleasure to be had in the heightened awarenesses arising
from resistance in performance, as well as letting go and being transported in its ow (cf. Do man 2019:
184).

Carey’s perspective brings the discussion back to a fundamental question that underpins performing the C24.P44
Concert, and Cage’s other works that adopt this approach: the extent to which it is possible to divorce one’s
attention from one’s fellow musicians. This is a question that trumpeter Jonathan Impett considered during
interview:

Now, to the best of my recollection it doesn’t say in the instructions to the trumpet part, ‘Don’t C24.P45
listen to other people’. Now this may well be a Cageian ethos which is entirely appropriate and
entirely substantiated by his various comments. I think what’s interesting performing the piece is
the degree to which it’s very di cult to switch that o , and you nd yourself playing with the urge
to play with other people. […] I think that challenges your listening, it challenges your
musicianship, it challenges your experience and your understanding of what it is to play with other
32
people.

p. 520 Impett’s statement identi es an important issue at the core of this chapter: there is a tension between the C24.P46
meso and the micro levels of time consciousness in a performance of the Concert, whereby an encultured
understanding of Cageian performance practices might be at odds with the intuitive sense of ensemble that
arises from sharing a space, time (in a broad, non-metrical sense), and bodily presence. Moreover, as
discussed, Cage’s instructions that each instrumental part can be conceived of as a ‘solo or a part in an
ensemble, symphony, or concerto’ invite a reconsideration of the various di erent roles and responsibilities
that the performer might take on within the group. Indeed, some of the performers’ comments suggest that
there is a sense that these roles were necessarily uid in the moment of performance, with the musicians’
elds of attention intersecting one another in sometimes quite unpredictable ways.

Conclusions C24.S8

While this case study focuses on only one of the almost limitless range of realization possibilities that the C24.P47
Concert a ords, the Apartment House musicians’ responses to Cage’s notation were carefully considered,
rich, and diverse, and evidence a variegated performance practice surrounding the Concert that contradicts
the conventional view of what playing Cage’s music might involve. What is more, the performers’
experiences suggest that notions of a uni ed, intersubjective communality in ensemble performance are
inadequate to account for the wide range of modes of consciousness exhibited during performance. And yet
it is clear that if the Concert is evaluated on the basis of being able to isolate the players from one another, it
is surely a failure. Perhaps a more interesting question, then, is not whether it is possible to ‘truly’ suppress
the intersubjectivity of performance, but what the consequences of attempting to suppress it might be, both
for the musicians’ immediate experiences of performance and, more broadly, for the meaning and
signi cance of playing (indeterminate) music.
As already touched on, and as has been identi ed by commentators for more than 50 years (see e.g. Adorno C24.P48
1961), this case study exposes a tension between freedom and constraint operating at various levels in
indeterminate music. Indeterminacy has been famously held up as a model of social democracy,
emancipating the performer from the hegemonies of the Western classical tradition, and threatening the
established hierarchy between composer, performer, and listener. For Donald Anderson (2011), it represents
a mode of ‘noisy’ resistance that has the capacity to ‘frustrate and scramble codes of oppression’ within ‘a
network of power relations’ (p. 20). Yet, discussions of indeterminacy have tended to focus on its essentially

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
compositional aspects, rather than on how the music functions or is experienced in performance. Indeed,
these various claimed freedoms have not necessarily played out in practice, with attitudes of performers and
composers painting a somewhat more mixed picture. As Piekut (2011) has shown in his account of the
resistance of orchestral musicians to Cage’s Atlas Eclipticalis, and in the paradox of Cage’s advocacy for
p. 521 Tudor’s ‘creative—even subversive—engagement with the score’ (Thomas 2013: 111), performance both
foregrounds and complicates notions of authority and agency in indeterminacy.

To pick up on Carey’s comments, how does a performance of the Concert di er from that of any other C24.P49
concert hall situation? The di erence lies in what ‘playing together’ means in the context of performing the
Concert: the ensemble shares a performance space, a conductor, and (at a very basic level) the objective of
starting and nishing at the same time. Yet the routes taken by the musicians to reach this goal are
deliberately disrupted. The temporal indeterminacy of the Concert is instigated by Cage’s notation, but it is
not simply written into the score. Analysis of the micro-social dynamics of the Apartment House
performance shows that it is in ected by various other factors, such as the musicians’ individual
performance choices, previous training and performance experiences, and their attitudes towards Cage’s
music, as well as the practical constraints of realizing the unstable techniques within the uctuating
e ective time set by the conductor. The players’ practical time consciousness is shaped at the meso level by
their pursuit of a shared sense of aesthetic goal that requires deliberate resistance to rhythmic and temporal
coherence at the micro level. At the macro level, these conditions recon gure the social ‘systems of
in uence’ (Ponchione-Bailey 2016) within the orchestra—in terms both of the relationship between the
conductor and musicians and of the internal relationships within the ensemble—resulting in an altered
temporal sociality.

In light of this discussion, it might be assumed that a performance of the Concert would present an C24.P50
embodiment of anarchy, with no relationship between the conductors’ gestures and musicians’ actions. And
yet, watching and listening to the lm of the performance, the musicians’ poker faces belie the separate
togetherness being exercised through their playing. The music exhibits shifts in collective energy,
particularly at points where there is a sudden change in speed and performers noticeably adjust their
playing in response to the movements of Sheen’s arms (see Iddon et al. 2019). Despite the apparent
indeterminacy of Cage’s notation, then, the performance outcome is not a static assemblage of isolated
sounds: there remains a relationship between the conductor and ensemble’s actions. The situation is
complicated further by the pianist, who simply orbits this constellation of activity without apparently
33
intersecting it. The audience, too, is conspicuous in its absence in this discussion, which has focused
largely on the production of indeterminacy rather than its reception. One perspective is put forward in a
review of the concert by Oliver Thurley (2018), who identi es moments of ‘building agitation’ (p. 91) as the
conductor’s rotations accelerated. For Thurley, there is a cohesion in the ‘collective chaos’ of the
performance, whereby ‘delicate, sustained tones passed back and forth across the ensemble […] even where
individual interpretations run isolated or even contrary to each other’ (p. 91). The interplay between
performers and audience and its role in the construction of meaning in indeterminacy is an area that invites
further scrutiny, and would surely bring further nuance to this discussion.

In sum, the role of the conductor in the Concert for Piano and Orchestra is much more than simply one of C24.P51
utility: on the one hand Sheen’s function was purely chronometric and one of tyrannical ‘time
management’; but on the other, the mutual detuning of conductor and ensemble resulted in (among other
p. 522 responses) surprise, confusion, anticipation, frustration, and playfulness in performance, with the
conductor, and in many cases the performers, taking pleasure in its agonistic sociality. Understood as a
fundamentally socially inscribed and creative practice (DeNora 2011; Do man 2019), practical time
consciousness thus plays an active role in the assemblage of attention and experiences that constitute
ensemble performance. In doing so, it invites a reconsideration of the socio-musical interactions that
characterize indeterminacy, and how the dynamics of such interactions in ect notions of liberty,

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
constraint, and discipline. This, in turn, might open up possibilities for a practically and socially oriented
understanding of what it means to perform indeterminate music.

Acknowledgements C24.P52

This research was supported by the Arts and Humanities Research Council (Project Reference: C24.P53
AH/M008444/1). The project team’s archival research at Northwestern University Library’s John Cage
Collection was funded by a John Cage Research Grant. I am grateful to the John Cage Trust, and to Laura
Kuhn personally, for the use of Cage’s unpublished correspondence. I wish to thank the musicians of
Apartment House for their participation in this research, and to Angela Guyton for creating the lms for the
project. Finally, thanks are due to Laura Anderson, Eric Clarke, Mark Do man, Martin Iddon, Philip
Thomas, and Toby Young for their insightful comments on earlier versions of this chapter.

Notes
1. See e.g. Atlas Eclipticalis (1961–2), Cheap Imitation (1970–72), the later Etcetera (1973; 1985) pieces, and the ʻNumber C24.N1
Piecesʼ for orchestra such as 1O1 (1988) and 108 (1991), which develop more radically some of the themes expressed in the
Concert, but which were beyond the scope of this chapter.

2. This chapter is an output of the Arts and Humanities Research Council project, ʻJohn Cage and the Concert for Piano and C24.N2
Orchestraʼ, led by Philip Thomas and Martin Iddon, and based at the Universities of Huddersfield and Leeds (Project
Reference: AH/M008444/1). The main project findings are published in Iddon and Thomas (2020) and on the project
website: <https://cageconcert.org/>.

3. See Floris Schuilingʼs Ch. 25 in this volume for a reconsideration of Schütz (1951) and its consequences for Cookʼs work. C24.N3

4. The ʻinside/outside the roomʼ dialectic and its role in creativity is set out in an earlier article by Clarke et al. (2016). C24.N4

5. The etymology of the term ʻconcertʼ is varied: in perhaps its broadest sense, it can indicate an ʻassemblyʼ of musicians C24.N5
performing together (Weber 2001); in French, a concert denotes an instrumental work comprising separate pièces (Sadie
2001); the word also has links to the Italian concertare (ʻto arrange, agree, get togetherʼ) and ʻconsortʼ, the early English
instrumental ensemble (Weber 2001).

6. Indeed, there are over 16,000 possible combinations of the instrumental parts alone (Iddon and Thomas 2020: 268). C24.N6

p. 523 7. Detailed comparisons of the instrumental parts can be found in Iddon and Thomas (2020: ch. 3). C24.N7

8. For detailed analyses of the conductorʼs part, see Iddon and Thomas (2020: ch. 5) and Iddon et al. (2019d). C24.N8

9. Other recordings of the Concert di er widely in length. The first performance (Cage et al. 1992a[1958]) has a duration of 23 C24.N9
minutes and 13 seconds, and most other recordings of the piece last between 20 and 30 minutes.

10. This is not to suggest that the Concert could be described as improvisation as in Goehrʼs ʻimprovisation extemporeʼ (i.e. C24.N10
whereby improvisation fulfils a solely creative function). Indeed, a year a er the pieceʼs premiere Cage seems to have
explicitly rejected any connection between performing the Concert and improvisation, stating in a letter to Peter Yates that
while ʻnaturally each performance is uniqueʼ, ʻno part is improvised. Each person simply does what he has to do and which
interpenetrates with all the other parts but not in a way which anyone has planned or must controlʼ (Kuhn 2016: 205).
However, he subsequently referred to jazz in describing the Concertʼs sound world (see Cage 1961a[1958]: 31; Kuhn 2016:
249). Moreover, the pieceʼs requirement for performers to respond in the moment to the unpredictable temporal shi s set
by the conductor is suggestive of improvisatory behaviour, as is the idea that the piece is formed of a combination of solos
rather than a cohesive ensemble such as an orchestra—an aspect compounded by the separation of the musicians in the
performance space (see Kim 2012 for a detailed account of Cage employing a similar approach when working with jazz
musicians).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
11. Source: John Cage Collection, Northwestern University Music Library, box 9, folder 1. Quotation used by permission of the C24.N11
John Cage Trust.

12. The spatial distribution of musicians was a common feature of later performances of the Concert (see Iddon and Thomas C24.N12
2020: ch. 5) and some of Cageʼs other works, the ʻhappeningsʼ in particular (Kirby 1965).

13. Accounts of the Concertʼs riotous premiere almost wholly rely on reception history and are likely to be apocryphal, or at C24.N13
least exaggerated. Moreover, the fact that Cage continued to work with a number of the musicians who were the ʻculpritsʼ
of the disruption suggests that his dissatisfaction with their playing was short-lived at most (see Iddon and Thomas 2020:
ch. 3).

14. See Iddon et al. (2019b) for a film of the performance. C24.N14

15. All interviews were conducted by the project team, filmed by Angela Guyton, and transcribed by Emily Payne. The C24.N15
members of Apartment House who participated in the project are Bridget Carey (Viola 1), Reiad Chibah (Viola 2), Andrew
Digby (Trombone), Ruth Ehrlich (Violin 3), Christian Forshaw (Bassoon and Baritone Saxophone), Jonathan Heilbron
(Double Bass), Jonathan Impett (Trumpet), Anton Lukoszevieze (Cello), Aisha Orazbayeva (Violin 1), Melvyn Poore (Tuba),
Nancy Ru er (Flute), Jack Sheen (Conductor), Hilary Sturt (Violin 2), Philip Thomas (Piano), and Vicky Wright (Clarinet).
Video recordings of the interviews are available on the accompanying website

16. Interview with Jack Sheen, 3 July 2017. C24.N16

17. Ibid. C24.N17

18. Interview with Anton Lukoszevieze, 4 July 2017. C24.N18

19. Interview with Reiad Chibah, 4 July 2017. C24.N19

20. Ibid. C24.N20

21. Interview with Jack Sheen, 21 May 2017. C24.N21

22. Interview with Jack Sheen, 3 July 2017. C24.N22

p. 524 23. Interview with Ruth Erlich, 3 July 2017. C24.N23

24. Interview with Jack Sheen, 3 July 2017. C24.N24

25. Ibid. C24.N25

26. While it is arguably the case that all conductors will be aware of how their performance is perceived by the audience, it is C24.N26
likely that their primary focus would be their relationship to the musicians on stage.

27. Interview with Hilary Sturt, 3 July 2017. C24.N27

28. Interview with Vicky Wright, 4 July 2017. C24.N28

29. The musiciansʼ (apparent lack of) responses to one another contrast strikingly to their behaviour during the other piece in C24.N29
the concert programme, Christian Wol ʼs Resistance (2016–17), which can be viewed at Iddon et al. (2019a).

30. Interview with Jack Sheen, 3 July 2017. C24.N30

31. Interview with Bridget Carey, 4 July 2017. C24.N31


32. Interview with Jonathan Impett, 3 July 2017. C24.N32

33. Thomas has discussed his approach to preparing and performing the Solo for Piano elsewhere. See Thomas (2013) and C24.N33
Iddon et al. (2019c).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
References C24.S9

Adorno, T. (1961). Vers une musique informelle. In Quasi una fantasia: Essays on modern music, 269–322. Verso. C24.P54
Google Scholar Google Preview WorldCat COPAC

Anderson, D. (2011). Power and indeterminacy: The noisy networks of Foucault, Cage, Burroughs, and Delaney. Doctoral C24.P55
dissertation, University of Seattle.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Apartment House (2017). Concert for Piano and Orchestra. On CC [CD]. Huddersfield Contemporary Records. C24.P56
Google Scholar Google Preview WorldCat COPAC

Barrett, G. D. (2016). A er sound: Toward a critical music. Bloomsbury. C24.P57


Google Scholar Google Preview WorldCat COPAC

Blum, D. (1987). The art of quartet playing: The Guarneri Quartet in conversation with David Blum. Gollancz. C24.P58
Google Scholar Google Preview WorldCat COPAC

Brooks, W. (2002). Music and society. In D. Nicholls (ed.), The Cambridge companion to John Cage, 214–226. Cambridge C24.P59
University Press.
Google Scholar Google Preview WorldCat COPAC

Cage, J. (1960[1958]). Concert for Piano and Orchestra. Peters Edition. C24.P60
Google Scholar Google Preview WorldCat COPAC

Cage, J. (1961a[1958]). Composition as process: Changes. In Silence: Lectures and writings, 18–34. Wesleyan University. C24.P61
Google Scholar Google Preview WorldCat COPAC

Cage, J. (1961b[c.1959]). Indeterminacy. In Silence: Lectures and writings, 35–40. Wesleyan University Press. C24.P62
Google Scholar Google Preview WorldCat COPAC

Cage, J., et al. (1992a[1958]). Concert for Piano and Orchestra. The 25-year retrospective concert of the music of John Cage. CD. C24.P63
Wergo.

Cage, J., et al. (1992b[1958]). Liner notes. In The 25-year retrospective concert of the music of John Cage. CD. Wergo. C24.P64

Cage, J., and Charles, D. (1981[1976]). For the birds (trans. R. Gardner, ed. T. Gora and J. Cage). Marion Boyars. C24.P65
Google Scholar Google Preview WorldCat COPAC

Cage, J., and Kostelanetz, R. (eds) (2000[1993]). John Cage, writer: Selected texts (2nd edn). Cooper Square. C24.P66
Google Scholar Google Preview WorldCat COPAC

Campana, D. (2001). As time passes. In D. Bernstein and C. Hatch (eds), Writings through John Cageʼs music, poetry, and art, 120– C24.P67
136. University of Chicago Press.
Google Scholar Google Preview WorldCat COPAC

p. 525 Clarke, E., Do man, M., and Timmers, R. (2016). Creativity, collaboration and development in Jeremy Thurlowʼs Ouija for Peter C24.P68
Sheppard Skærved. Journal of the Royal Musical Association 141(1): 113–165.

Cline, D. (2019). Two concepts of indeterminacy in music. Musical Quarterly 102(1): 82–110. C24.P69
Google Scholar WorldCat

Cook, N. (2007[2004]). Making music together, or improvisation and its others. In Music, performance, meaning: Selected essays, C24.P70
321–341. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Cottrell, S. (2017). The creative work of large ensembles. In J. Rink, H. Gaunt, and A. Williamon (eds), Musicians in the making: C24.P71
Pathways to creative performance, 186–205. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Csíkszentmihályi, M. (1996). Creativity: Flow and the psychology of discovery and invention. HarperCollins. C24.P72

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

DeNora, T. (2011). Practical consciousness and social relation in MusEcological perspective. In D. Clarke and E. Clarke (eds), Music C24.P73
and consciousness: Philosophical, psychological, and cultural perspective, 309–325. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Dobson, M. C., and Gaunt, H. F. (2015). Musical and social communication in expert orchestral performance. Psychology of Music C24.P74
43(1): 24–42.
Google Scholar WorldCat

Do man, M. (2019). Practical time consciousness in musical performance. In R. Herbert, D. Clarke, and E. Clarke (eds), Music and C24.P75
consciousness 2, 170–186. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Faulkner, R. R. (1973). Orchestra interaction: Some features of communication and authority in an artistic organization. C24.P76
Sociological Quarterly 14(2): 147–157.
Google Scholar WorldCat

Goehr, L. (2016). Improvising impromptu, or, what to do with a broken string. In G. Lewis and B. Piekut (eds), The Oxford C24.P77
handbook of critical improvisation studies, vol. 1, 458–480. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Hø ding, S. (2019). A phenomenology of musical absorption. Palgrave Macmillan. C24.P78


Google Scholar Google Preview WorldCat COPAC

Huron, D. (2006). Sweet anticipation. MIT Press. C24.P79


Google Scholar Google Preview WorldCat COPAC

Iddon, M. (2013). John Cage and David Tudor: Correspondence on interpretation and performance. Cambridge University Press. C24.P80
Google Scholar Google Preview WorldCat COPAC

Iddon, M., Payne, E., Melen, C., and Thomas, P. (2019a). Apartment House performs Christian Wol ʼs Resistance. John Cage and C24.P81
the Concert for Piano and Orchestra. https://cageconcert.org/project/outputs-and-activities/apartment-house-perform-christian-
wol s-resistance/>.
WorldCat

Iddon, M., Payne, E., Melen, C., and Thomas, P. (2019b). Apartment House and Philip Thomas perform the Concert for Piano and C24.P82
Orchestra. John Cage and the Concert for Piano and Orchestra. https://cageconcert.org/performing-the-concert/apartment-
house-and-philip-thomas-perform-the-concert-for-piano-and-orchestra/>.
WorldCat

Iddon, M., Payne, E., Melen, C., and Thomas, P. (2019c). Philip Thomas: Performing the Solo for Piano. John Cage and the Concert C24.P83
for Piano and Orchestra. https://cageconcert.org/performing-the-concert/solo-for-piano/>.
WorldCat

Iddon, M., Payne, E., Melen, C., and Thomas, P. (2019d). The conductor. John Cage and the Concert for Piano and Orchestra. C24.P84
https://cageconcert.org/performing-the-concert/the-conductor/>.
WorldCat

Iddon, M., Payne, E., and Thomas, P. (2019). Disruption and discipline: Approaches to performing John Cageʼs Concert for Piano C24.P85
and Orchestra. Music + Practice 5. https://www.musicandpractice.org/volume-5/disruption-and-discipline-approaches-to-
performing-john-cages-concert-for-piano-and-orchestra/>.
WorldCat

Iddon, M., and Thomas, P. (2020). John Cageʼs Concert for Piano and Orchestra. Oxford University Press. C24.P86

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

p. 526 Kim, R. (2012). John Cage in separate togetherness with jazz. Contemporary Music Review 31(1): 63–89. C24.P87
Google Scholar WorldCat

Kirby, M. (1965). The new theatre. Tulane Drama Review 10(2): 23–43. C24.P88
Google Scholar WorldCat

Kostelanetz, R. (2003[1987]). Conversing with Cage (2nd edn). Routledge. C24.P89


Google Scholar Google Preview WorldCat COPAC

Kotik, P. (1993). Liner notes. In: John Cage: Concert for Piano and Orchestra/Atlas Eclipticalis. CD. Wergo. C24.P90
Google Scholar Google Preview WorldCat COPAC

Kuhn, L. (ed.). (2016). The selected letters of John Cage. Wesleyan University Press. C24.P91
Google Scholar Google Preview WorldCat COPAC

Piekut, B. (2011). Experimentalism otherwise: The New York avant-garde and its limits. University of California Press. C24.P92
Google Scholar Google Preview WorldCat COPAC

Ponchione-Bailey, C. (2016). Tracking authorship and creativity in orchestral performance Doctoral dissertation, University of C24.P93
Oxford.
Google Scholar Google Preview WorldCat COPAC

Pritchett, J. (1993). The music of John Cage. Cambridge University Press. C24.P94
Google Scholar Google Preview WorldCat COPAC

Pritchett, J. (2004). David Tudor as composer/performer in Cageʼs Variations II. Leonardo Music Journal 14: 11–16. C24.P95

Sadie, J. (2001). Concert (i). Grove Music Online. C24.P96


https://www.oxfordmusiconline.com/grovemusic/view/10.1093/gmo/9781561592630.001.0001/omo-9781561592630-e-
0000006239>.
Google Scholar Google Preview WorldCat COPAC

Sawyer, R. (2003). Group creativity: Music, theater, collaboration. Erlbaum. C24.P97


Google Scholar Google Preview WorldCat COPAC

Sawyer, R. (2014). Musical performance as collaborative practice. In M. S. Barrett (ed.), Collaborative creative thought and C24.P98
practice in music, 271–286. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Schütz, A. (1951). Making music together: A study in social relationship. Social Research 18(1): 76–97. C24.P99
Google Scholar WorldCat

Seddon, F. (2004). Empathetic creativity: The product of empathetic attunement. In D. Miell and K. Littleton (eds), Collaborative C24.P100
creativity: Contemporary perspectives, 65–78. Free Association Books.
Google Scholar Google Preview WorldCat COPAC

Seddon, F., and Biasutti, M. (2009). Modes of communication between members of a string quartet. Small Group Research 40(2): C24.P101
115–137.
Google Scholar WorldCat

Thomas, P. (2013). Understanding indeterminate music through performance: Cageʼs Solo for Piano. Twentieth-Century Music C24.P102
10(1): 91–113.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471316 by National Science & Technology Library user on 26 May 2023
Thurley, O. (2018). Apartment House: Wol , Cage, ʻPerforming Indeterminacyʼ. Tempo 72(283): 90–92. C24.P103
Google Scholar WorldCat

Valkenburg, J. (2010). From bars to inches (to seconds): Timekeeping in the music of John Cage. Tijdschri voor Muziektheorie C24.P104
15(1): 68–75.
Google Scholar WorldCat

Waddington, C. (2017). When it clicks: Co-performer empathy in expert ensemble playing. In E. King and C. Waddington (eds), C24.P105
Music and empathy, 230–247. Routledge.
Google Scholar Google Preview WorldCat COPAC

Weber, W. (2001). Concert (ii). Grove Music Online. C24.P106


https://www.oxfordmusiconline.com/grovemusic/view/10.1093/gmo/9781561592630.001.0001/omo-9781561592630-e-
0000006240>.
Google Scholar Google Preview WorldCat COPAC

Weeks, P. (1990). Musical time as a practical accomplishment: A change in tempo. Human Studies 13(4): 323–359. C24.P107
Google Scholar WorldCat

Weeks, P. (1996). Synchrony lost, synchrony regained: The achievement of musical coordination. Human Studies 9(2): 199–228. C24.P108
Google Scholar WorldCat

Wenger, E. (1998). Communities of practice: Learning, meaning, and identity. Cambridge University Press. C24.P109
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
CHAPTER

25 ‘Making, Not Filling Time’: TimeC25and Notation in


Improvised Musical Performance 
Floris Schuiling

https://doi.org/10.1093/oxfordhb/9780190947279.013.28 Pages 527–548


Published: 08 December 2021

Abstract
Improvisation and notation are frequently opposed in terms of transience versus permanence, an
opposition that re ects broader Eurocentric ideas of orality and literacy. Confronting such binary
distinctions, Schuiling describes the use of notation by three groups of improvising musicians,
showing how notations mediate their understanding of time. This forms the basis of a critique of
Alfred Schütz’s in uential account of social interaction in musical performance. Schuiling argues that
Schütz’s distinction of an ‘inner time’ of music and an ‘outer time’ mapped by the score remains tied
to a work-centred musical ontology, and fails to attend to the making of time in the course of
performance. Drawing on his eldwork, Schuiling reconsiders the work of Maurice Halbwachs, the
primary target of Schütz’s argument. Rather than understanding music as an object of inner
contemplation, Halbwachs provides a view of music and temporality as a way of opening up to the
world.

Keywords: improvisation, notation, time, Alfred Schütz, Maurice Halbwachs, musical performance
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

IMPROVISATION and the performance of composed music have often been distinguished on the basis of their C25.P1
di erent relations to time. Especially in the twentieth century, when improvisation was mostly associated
with jazz and African American culture, while composition became an increasingly academic a air, this
distinction was mapped onto a contrast between oral and literate culture. While the older notion of
‘extemporization’ places improvisation outside time, modern writers have often argued or assumed that
improvisation a ords a truer, more authentic relation to temporal experience. We can see this, for instance,
in Ben Sidran’s Black talk, in which he describes an aesthetics of Black musical performance as an alternative
to European narratives and discourses:
Oral cultures use only the spoken word and its oral derivatives, i.e., musical representations of C25.P2
basic vocalizations. This determines the referents of oral perception in general. […] Whereas paper
and ink are the medium of the literate man, oral communication is immediate. That is, oral
communication is free from intervention of a medium. It is a ‘direct presence.’ To paraphrase
McLuhan, the message is the medium. The oral man thus has a unique approach to the
phenomenon of time in general: he is forced to behave in a spontaneous manner, to act and react
(instantaneous feedback) simultaneously. As a consequence of this perceptual orientation, oral

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
man is, at all times, emotionally involved in, as opposed to intellectually detached from, his
environment through the acts of communication. This can be called the basic actionality of the oral
personality. McLuhan has characterized this lack of intellectual detachment as contributing to a
superior sense of community, a heightened ‘collective unconscious’ and ‘collective awareness,’
within oral cultures. It is su cient for the moment to suggest that sound is eeting and one must
react immediately or lose the perceptual experience entirely.

(Sidran 1995: 2–3)

p. 528 Sidran’s text is shot through with binaries. Black and White culture are opposed through binaries of orality C25.P3
versus literacy, time versus space, immediacy versus technology, spontaneity versus detachment, emotion
versus intellect, communality versus individualism, practice versus theory.

As Sidran’s references to media scholar Marshall McLuhan indicate, this binary oppositional discourse is C25.P4
part of much broader intellectual trends in twentieth-century thought, in which modern technology, and
particularly writing and printing, is seen to be structured around the eye rather than the ear. Thus, in the
words of McLuhan’s student Walter Ong, orality is ‘empathetic and participatory rather than objectively
distanced’ (Ong 2005: 45), and literacy thus separates and alienates people from the communality and
‘interiority’ of sound (p. 69 .). Jonathan Sterne has called this logic the ‘audio-visual litany’: the idea that
modernity has focused on the eye (considered as the sense of detachment and individualism) at the expense
of the ear (considered as the sense of communion and authenticity) (Sterne 2003: 15). In Sterne’s
formulation, vision is frequently associated with space, while hearing is associated with time. However, a
more precise distinction might be made: modernity is associated with linear, measurable time, while its
Others are said to have a more cyclical or static notion of time. In twentieth-century anthropology, studies
abound that contrast the static or cyclical time-consciousness of non-Western societies to the linearity of
1
time in the modern West (some of the most in uential include Bourdieu 1963; Geertz 1976; Whorf 1964).

The consciousness of time and the temporality of consciousness have also been a central theme in C25.P5
philosophy, with musical experience frequently being the primary object of re ection. Some philosophers,
most notably Henri Bergson and Martin Heidegger, have argued that the subjective experience of duration is
the truest form of temporal reality. Bergson’s Time and Free Will argued that the only reality is the present,
and that the measurement of time amounts to a confusion of time and space; instead of such a
‘spatialization’ of time, freedom consists of a return to ‘pure duration’ (Bergson 2001). The nal pages of
Heidegger’s Being and time somewhat similarly argue that ‘measured time’ is a result of human beings’
‘thrownness’, and that anyone who lets their existence be governed by clocks or calendars leads an
inauthentic life, forgetting the ‘making present’ that is characteristic of authentic temporality (Heidegger
1962: 457 .).

This distinction between a lived reality and measured abstraction of time, as we saw in the quotation by C25.P6
Sidran, has also in uenced writing on music, where improvisation has been described as a practice carried
out entirely ‘in the moment’, and hence is regarded as more authentic, expressive, participatory, and
creative. A musical composition, conversely, is considered as a timeless, abstract idea, and musicians
performing a composition are frequently thought simply to follow the linear order of notes as established in
the score. Such absolute distinctions, however, are generally no longer upheld: recent scholarship on
musical performance has overwhelmingly argued that musical meaning is not inherent to the work, but
created in performance, and that the complex processes of timing, mutual coordination, embodied
knowledge, and social interaction involved in this creation cannot be accounted for by assuming that the
music is simply ‘in’ the score (see e.g. Abbate 2004; Clarke and Do man 2017; Cook 2013; Doğantan-Dack
p. 529 2011; Rink 1990; Taruskin 1995). In jazz and improvisation studies, meanwhile, scholars are increasingly
moving away from a conception of improvisation in terms of immediacy, investigating instead how the
learning and performing of jazz and improvised music are socially and technologically mediated (see e.g.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
Monson 1996; Prouty 2006; Tackley 2010; Schuiling 2019).

This chapter makes a contribution to these complementary movements by examining the relation of music C25.P7
notation to the temporality of performance, focusing speci cally on the role of notation in improvised
music. Contrary to popular belief, improvising musicians frequently use various forms of notated music and
other kinds of compositional frameworks or models in their creative practice. I draw on case studies from
eldwork with three groups of improvisers based in the Netherlands that use di erent forms of notation
(understood broadly as the representation of musical ideas through signs or symbols) in their work. The
rst group is the Instant Composers Pool (ICP) Orchestra, a group that played a key role in the development
of improvised music in the Netherlands (Schuiling 2019). The second is a group of students at HKU
Conservatory in Utrecht, who learn Kobranie, a system of conducted improvisation developed by their
teacher, Esmée Olthuis. The third is the Genetic Choir, a semi-professional improvising choir whose artistic
director, Thomas Johannsen, aims to identify processes and strategies of improvising, and feeds these back
2
into their practice to guide new improvised pieces. Together, these three examples cover a broad spectrum
of uses of notation, with the ICP Orchestra using a large repertoire of notated compositions, Kobranie being
a standardized system of gestures and hand signals that may be employed di erently in each performance,
and the Genetic choir using only certain rules, strategies, or concepts that do not take a material form, to
guide the improvisers’ actions. There are some important di erences between these three groups and their
uses of notation as well as their conceptions of musical creativity. My main concern in this chapter,
however, is not ethnographical but theoretical. My description of their practices thus has a heuristic rather
than an illustrative or corroborative goal, and draws on my experiences, observations, and conversations
with these musicians to argue for a conception of temporality in musical performance that is more
di erentiated than the usual binaries allow. Such a conception may contribute to a music scholarship that is
sensitive to the wide variety of notation systems in use today, and that sees the writing and reading of
notation as integral to the creative and collaborative processes of musical performance.

The Temporality of Musical Performance C25.S1

Perhaps the most in uential argument concerning the temporality of musical performance, particularly C25.P8
with regard to the comparison between improvised and composed forms of music, is found in philosopher
p. 530 of sociology Alfred Schütz’s ‘Making Music Together’ (1951). Long forgotten, at least by musicologists, it
was brought back to their attention by Nicholas Cook in one of his arguments for a musicology of
performance (Cook 2007[2004]; see also 2013: 236–238). Schütz was best known for his phenomenology of
social reality. ‘Making Music Together’ re ects his interest in intersubjectivity, as well as his background in
early twentieth-century Viennese intellectual life, taking musical performance (and particularly the string
quartet) as an exemplary case study for what it means to be in a social relationship.

Schütz turns to music because, as he states in his opening sentence, ‘music is a meaningful context which is C25.P9
not bound to a conceptual scheme’ (1951: 76). The performance of music involves a set of complex social
relations between composer, performer(s), and listener(s), which work together in a way that can be
characterized as a form of communication. He argues that the sociality and substance of this process,
however, do not rest on a set of mutually held concepts or a common language shared by all persons
involved, as the meaning of music is not primarily linguistic. This argument is made explicitly not to better
understand music speci cally, but to make a point about social relations in general, namely to ask the
question:

whether the communicative process is really the foundation of all possible social relationships, or C25.P10
whether, on the contrary, all communication presupposes the existence of some kind of social
interaction which, though it is an indispensable condition of all possible communication, does not

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
enter the communicative process and is not capable of being grasped by it.

(Schütz 1951: 78)

Schütz uses the example of musicians making music together to argue for the latter: the phenomenological C25.P11
position that subjectivity is not the result of language or belonging to a society, but of a more fundamental,
pre-linguistic being part of the world.

In making this argument, Schütz turns against his contemporary Maurice Halbwachs, a sociologist best C25.P12
remembered for his work on collective memory. Halbwachs had also used the example of musical
performance to highlight aspects of social interaction, arguing that (to paraphrase Schütz’s summary of his
work) since professional musicians can read notation, they are able to remember the music ‘itself’, while
amateurs (or more speci cally non-readers of notation) remember music only by virtue of association of
the music with words (such as the lyrics to a melody) or with bodily movement (such as dances or marches).
Schütz argues that Halbwachs mistakenly identi es musical thought with its communication,
communication with musical language, musical language with music notation, and notation with the social
background of musicians (Schütz 1951: 82).

By contrast, Schütz de nes music as ‘a meaningful arrangement of tones in inner time’, referring to C25.P13
Bergson’s durée as music’s primary site of existence:

The ux of tones unrolling in inner time is an arrangement meaningful to both the composer and C25.P14
p. 531 the beholder, because and in so far as it evokes in the stream of consciousness participating in it
an interplay of recollections, retentions, protentions, and anticipations which interrelate the
successive elements.

(Schütz 1951: 88)

This ‘inner time’ is contrasted with the ‘outer’ time, measurable by clocks and represented by music C25.P15
notation. Outer time is completely and categorically distinct from musical reality; musical meaning is
essentially polythetic (experienced step by step, moment by moment as an unfolding in inner time) and
cannot be grasped monothetically—as one might, perhaps, with a letter or a book; after having read it
polythetically one can grasp it as a whole, no longer needing to read it step by step again (pp. 90–91).
Musicians making music together act as an intermediary, participating in the inner time of both composer
and listener (p. 93) and sharing in a ‘mutual tuning-in relationship’ that is constitutive of intersubjectivity,
the experience of togetherness as a ‘We’ (p. 96). Thus, a shared temporality is argued to form the basis for
the intersubjectivity that forms a condition for social life, and this is exempli ed by music.

Cook (2007) draws on Schütz’s argument as a basis for developing a musicology of performance, and to C25.P16
formulate alternative ways of understanding performance to the paradigm of ‘reproduction’ that a work- or
text-based understanding of music implies. Part of his argument is to interrogate the distinction between
classical music as a literate tradition and jazz improvisation as an oral tradition (mentioning Sidran’s work
as one example of how this distinction is made). Surveying the research on the close interaction between
musicians that drives the process of jazz improvisation, Cook argues that such social and creative dynamics
are equally part of the performance of Western art music, and turns to Schütz for his argument that ‘there is
no di erence in principle between the performance of a string quartet and the improvisations at a jam
session of accomplished jazz players’ (Schütz 1951: 96). In both cases (and in fact in all musical
performance), the musicians engage in close collaboration and make precise adjustments of their
movements and timing in response to each other’s contributions. They share an inner time, and are
involved in a ‘mutual tuning-in’. To think that this is less the case in a performance tradition with written
music is to confuse the ‘literary’ appreciation of such written music with the reality of performance, and to
assume ‘that performance means bringing out something that is already there in the score, composed into it

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
and just waiting to be released by the performer’ (Cook 2007: 338).

Cook’s advocacy for the performative turn in music scholarship has been very in uential, and so has his C25.P17
argument that the concept of musical performance as a reproduction of what is ‘in’ the score misconstrues
the processes of performance, the social interaction between musicians, and the skill inherent to making
music. However, as I shall argue, Schütz’s account of the temporality of musical performance has some
signi cant limitations, and it is in fact an inadequate basis for understanding the interaction between
musicians. My argument was initially stimulated by a remark made by ICP cellist Tristan Honsinger. He
described to me, with some di culty in trying to formulate exactly what he meant, how improvised music
p. 532 could attain what he called ‘a relative content’, referring to his collaborations with the ICP’s leader Misha
Mengelberg as well as with Derek Bailey:

There was a time where I saw…I think this is possible, [to make music] without any indication, no C25.P18
power, no hierarchy. So I started when I was in my twenties, and what I was doing, I was lling up
the time. From this time to that time…and in working with Derek Bailey, I realized that he is not
lling time, he’s making time. And I realized yes this is it, this is partly it. And then I learned from
Misha the need to be very clear, and to take away instead of to add […]. So in order to make
something richer, you take away. These are two very important secrets in order to get to a ow of
3
content. That’s all I can really put into words.

Where Schütz distinguished, along with Bergson, between an inner and an outer time, assigning musical C25.P19
reality only to the former, Honsinger makes a distinction between making time and lling a time that has
already been made; this suggests that the temporality of musical performance emerges from the interaction
on stage, and thus emphasizes the creative achievement of performance.

Schütz’s two modes of temporality come ready-made: there is either the measurable and regular ‘outer C25.P20
time’ of material reality, or the intersubjective inner time that a listener ‘taps into’ as though tuning to a
radio station. His account of musicians coordinating their movements in performance, which is so central to
Cook’s argument, is unconvincing as a result. Schütz writes that performers navigate a ‘pluridimensionality
of time’, executing actions in the external world while also tuned into the inner time in which the music
unfolds. He describes this as follows: ‘The coperformers […] have to execute activities gearing into the outer
world and thus occurring in spatialized outer time. Consequently, each coperformer’s action is oriented not
only by the composer’s thought and his relationship to the audience but also reciprocally by the experiences
in inner and outer time of his fellow performer’ (Schütz 1951: 94). Thus, according to Schütz, in order to
make music, musicians are constantly trying to merge two temporalities that are fundamentally
incommensurable; moreover, the playing of their instruments, as it occurs in ‘outer time’, is essentially
unmusical.

If anyone ‘makes’ musical time in Schütz’s account, it is the composer, and the ‘mutual tuning-in C25.P21
relationship’ that he describes occurs between composer and listener—the listener tapping into (as it were)
the ow established by the composer, ‘performing with him step by step the ongoing articulation of his
musical thought’ (1951: 90). This is exactly the Platonic ideal of ‘a music that travels instantaneously from
mind to mind’ that Cook has criticized as part of his arguments for a musicology of performance, as it
makes the performance of music secondary to its existence (Cook 2013: 9). Far from rehabilitating the
performer as a creative agent, Schütz remains rmly within the paradigm of reproduction, as the creation of
music as ‘a meaningful arrangement of tones in inner time’ is completely the work of the composer, and not
the performers’ creative achievement.

p. 533 Schütz’s distinction of an aural inner time as authentically (and immediately) social from a visual, C25.P22
spatialized outer time is a clear instance of Sterne’s ‘audio-visual litany’ (2003: 15). What is odd about
Schütz’s version is that, unlike Bergson, for instance, rather than arguing that the idea of measurable time

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
is a confusion of time and space, Schütz seems to assume its reality and its accuracy as a description of
material existence—accepting a Newtonian description of time as a uniform line. He a rms that ‘outer
time is measurable’ and describes it as ‘the time that the musician “counts” in order to assure the correct
“tempo” ’ (1951: 89). Inner time, as in Bergson’s understanding, is never described except in negative terms
—as the opposite of outer time. Although at the end of his article Schütz announces that ‘the theory of the
tuning-in relationship’ (1951: 97) will be explored further in a future paper, he never returned to this topic.

Recent work on time in philosophy and the social sciences has largely moved away from a Bergsonian C25.P23
conception of subjective duration. We might see Bergson’s emphasis on the primacy of subjective duration
as what Bernard Stiegler has diagnosed as the ‘discourse of the fall’ inherent in much thinking about
temporality and human origins. This discourse (which he associates with Rousseau and Plato) posits a true
experience of time by the human soul in the realm of Ideas or a romanticized Nature, which is distorted by a
‘fall’ into the body, into technology, and into history (Stiegler 1998: 96–97). Like Sidran, whose work I
discussed at the start of this chapter, it posits an original, authentic, and immediate experience of time that
is led astray by the confusions of modernity. The discourse, however, fails to truly account for temporality,
because ‘there is never anything, at the origin, but the fall outside it’ (p. 101). Stiegler is one of various
scholars who have recently viewed techniques and technologies not as distortions of a ‘true’ primordial
temporality, but precisely as strategies for creating and negotiating time in everyday practices, constituting
temporalities that are at once technical, epistemic, and ethical (Bear 2016; Birth 2012; Born 2015; Munn
1992).

From this perspective, we might argue that notation, like other forms of technology, does not simply belong C25.P24
to a uniform ‘external time’ but constructs multiple temporalities that are not easily classi ed as ‘inner’ or
‘outer’. As Mark Abel has argued (2014), phenomenological theories of time as duration—and Schütz’s in
particular—seem to consider music mostly in terms of melody and harmony, at the expense of rhythms and
grooves. Indeed, the scare quotes in Schütz’s description of a musician ‘counting’ in a ‘tempo’ suggest as
much. Another description of outer time is that ‘which can be measured by counting or the metronome or
the beat of the conductor’s baton’ (Schütz 1951: 95). However, the purpose of a metronome (or the beat of a
conductor) is not primarily to measure time but to establish a rhythm, which can be used as a reference for
the mutual coordination of musicians’ movements to make music—one might say it is a technology for
making rather than measuring time. In fact, the measurement of time is perhaps better understood not as a
‘spatialization’ but as essentially a calibration of rhythms. Music notation maps out the mutual
coordination of musicians just as our calendars account for the rhythmic movement of the planet, and our
clocks divide this rhythm into smaller units.

p. 534 Hence, contrary to Bergson’s proto-phenomenological framework, time is not given solely by the human C25.P25
subject, but emerges from our interaction with our material environment. Musicological understandings of
performance and creativity have recently similarly moved away from an understanding of music as a purely
ideal phenomenon to address its emergence from the interaction between humans and musical technologies
and material objects (Born 2005; Burnard 2012; Clarke and Do man 2017; Clarke et al. 2013; Cook 2018;
Dolan 2013; Payne 2016). Following this movement to regard music as a ‘paradigmatic multiply-mediated’
form of creative practice (Born 2005: 7), the question of how we might reconsider the temporality of music
is signi cant, particularly with regard to the relation between text and performance, which is so central to
Schütz and Cook’s arguments.
The creative work of musicians, I suggest, is not adequately accounted for by a binary distinction between C25.P26
temporalities of either a uniform ow or a measurable series of evenly-spaced moments. The concept of
‘ ow’, also mentioned by Honsinger when describing his approach to improvisation, is often understood as
a psychological state of consciousness (most prominently as explored in the work of Csíkszentmihályi
2008). This concept came up in various ways during my eldwork, but it was used not to refer to an inner
mental state, or even a shared mental state, as in Keith Sawyer’s concept of ‘group ow’ (Sawyer 2017), but
rather to describe a uid sense of interaction with sounds, instruments, and fellow musicians—what

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
anthropologist Tim Ingold might call a correspondence with the ow of materials (Ingold 2010; 2012; 2017).
Rather than the idea of an unmediated inner time that dreams of a pure presence before all representation,
we might more fruitfully inquire into how musical media, including notation, con gure such interactions.
As I shall argue at the end of this chapter, the work of Halbwachs may in fact be a good starting point for
such inquiry.

Notations in Improvised Music C25.S2

The ICP Orchestra C25.S3

Honsinger’s remark about ‘making time’ was all the more signi cant because the ICP Orchestra performs a C25.P27
large repertoire of compositions, of which most were written by its leader and pianist, Mengelberg.
Performances by the group constitute a kind of improvised collage of a selection of these pieces, which
range from fully notated compositions to mobile forms and from jazz lead sheets to graphic scores. Before a
performance, the group makes a set list containing some compositions, as well as identifying a few groups
of three or four musicians of the orchestra (which consists of ten people) to play free improvisations. In the
performance, these items are ideally played without stopping, improvising transitions from piece to piece,
to a small group improvisation, to another composition, and so on (Schuiling 2019a).

p. 535 In addition, far from an ideal of improvisation as a collective musical ow, the group cultivates a form of C25.P28
interaction that allows for disagreement and forms of musical pestering in performance (Adlington 2013:
97–136). Mengelberg and drummer Han Bennink, the two founding members that were still part of the
group during my eldwork, were well known for their antagonistic duo performances, where they often
tried to undermine each other’s playing. Bennink, when discussing a recording I had made of an
improvisation by him, violinist Mary Oliver, and trombonist Wolter Wierbos, made clear that this way of
working still informs his practice today. He commented on his playing in the middle of the improvisation:
‘I’m working towards an ending, you can hear that clearly. Those dynamics, less, less, softer. But because
4
I’ve been playing so loudly Mary would like to go on for a while, while I keep trying to make an ending.’ As
Bennink stopped playing, Oliver took her chance to play some more now that she was audible. After a few
moments, Bennink attempted to sabotage the music by playing a loud crash on his cymbals. He commented:
‘Now I don’t think it’s on the level as where I wanted to end it. You see how sel sh this is? It’s limping along
without the quality that it had when I was still playing, and I was working very hard towards that ending.’
When I asked him whether he was actually just trying to get his way, he replied: ‘Yes, of course! I’m trying
to be in charge.’

Though Bennink might represent a somewhat extreme case, his comments suggest that his experience of C25.P29
temporality as an improviser is not one of tuning into a musical ow, but rather one that emerges from the
constant negotiations in the interactions and disagreements with his fellow performers. How do notations
t into this process? With the variety of compositions in the ICP’s repertoire, its pieces might be considered
as constructing di erent temporalities. Some of the more classically inspired and fully notated
compositions establish the experience of intensity, of speeding up, slowing down, of climax and stasis of the
kind that Schütz seemed to have in mind in his description of music as an arrangement of tones in inner
time. The more jazz-in uenced pieces are more about establishing a groove, which may be divided into
sections such as introductions, backgrounds, transitions, or codas, with a cyclical return of choruses over
which to play solos. Mobile form pieces establish a more dislocated temporality, with fragments that
suggest being part of a larger whole played in unconnected or detached ways, perhaps with improvised
sections in between.

In each case, however, such temporalities are the result of the creative e orts of the musicians and their C25.P30

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
mutual coordination, and are not simply given by the score. In fact, tenor saxophonist Tobias Delius
explained that the pieces in their repertoire are frequently used to interrupt the musical ow:

Many people say that improvisation can be too chaotic and then there is the ‘guiding hand’ of the C25.P31
composer or a piece to bring some sense of structure, but I think it’s the other way around. The
purpose of the written material is to disrupt a ‘nice ow’ of improvisation. It can create more
5
anarchy than improvisation sometimes. […] The compositions play their own part.

p. 536 The e ect of such interruptions of an improvisation (or of another piece) is to emphasize that the music is a C25.P32
result of the interaction and mutual coordination of the musicians, even when they play from a score. In
fact, the ICP has various ways in which they improvise by playing ideas found in their sheet music. This is
most obvious in the group’s improvised transitions between pieces or between a group improvisation and
the start of a piece. Such transitions might happen quite quickly, when someone cues the next piece and the
others join straight in, but usually these transitions are more elaborate. Musicians will play motives or other
ideas from the next piece, or combine ideas from the two pieces between which they are making a transition.
They may also physically point out aspects of the score to other musicians, that will then be interpreted in
di erent ways: a short melody might be played backwards or harmonized on the spot; or more radically,
they might point to a sign such as a bar line, a rest, or a clef and interpret it as a graphic score. In this way,
these transitions attain a ‘compositional’ quality in their own right, and the line between a piece and the
improvised transition that leads up to it and contains various anticipations of its musical ideas becomes
blurred. The notations thus do not stand at a remove from a ‘real’ temporality of musical performance, but
are tools for making time together.

Kobranie C25.S4

Although Kobranie is geared much more towards collective collaboration than the antagonistic C25.P33
performances of the ICP, it equally shows that improvisation is partly a matter of negotiating multiple
temporalities, and its use of notation is a means of doing so. As a form of conducted improvisation, one
person stands in front of a group of musicians and conducts them, using a system of around 100 signs to
communicate musical ideas. Since the 1970s such systems for conducted improvisation, including Butch
Morris’s ‘conduction’ and Walter Thompson’s ‘soundpainting’, emerged across di erent scenes of free
improvisers, and Kobranie was initially inspired by John Zorn’s Cobra (1984), although it quickly developed
6
in quite a di erent direction. Its signs include ‘start’, ‘stop’, ‘long note’, ‘short note’, ‘louder’, ‘softer’,
‘faster’, ‘slower’, ‘imitate this’, ‘support this’, ‘play a rhythm’, ‘play chords’, ‘play a melody’, ‘add a second
voice’, ‘give me a new idea’, ‘take a solo’, ‘tell me a story’, ‘record this’, ‘take it to the bridge’, ‘return’,
‘return to what we recorded’, ‘come together’, ‘come apart’, and ‘chaos’. The conductor (called a
‘processor’) points to a musician or group of musicians, gives a sign, and then cues its execution (though
some signs do not require this last cue). Olthuis teaches this system, which she developed in her own
practice as an improvising musician, to the rst-year students of one of the programmes at HKU Utrecht
Conservatory.

The signs and gestures of Kobranie do not specify a ‘piece’ beforehand, but adapt to the changing musical C25.P34
situation—somewhat like ICP musicians adapting their scores in their improvisations. The students
learning this system over the course of their rst year nonetheless have to learn to create a sense of shape or
form in the music that they are conducting (or ‘processing’). Their rst pieces are frequently rather
p. 537 fragmentary, consisting of short individual ideas or grooves, juxtaposed in rather disjunct ways and
without a strong sense of coherence. As they improve, they learn to ‘feed’ a musical situation, making it
clearer, intensifying its energy by adding more players or increasing the volume, and creating a sense of
climax or of conclusion. Towards the end of their rst year, the students can usually create a piece with
several ‘sections’, quickly organizing the musical situation without long introductions or transitions, and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
featuring di erent musical styles, instrumentations, and/or soloists.

Only a few signs let the processors control the form of the improvisation directly. The ‘bridge’ sign asks C25.P35
players to switch to a contrasting section on cue—a gesture that works remarkably well in practice, as
players may stop playing, include more rhythmic breaks, and move to subdominant or mediant harmonies
—and the processor can then direct a ‘return’ with a di erent sign. The ‘record’ sign has a similar function,
asking players to remember a particular musical situation and to return to it on cue. Other than these, all
shaping of the musical form is done by adding or removing players to and from the music, creating a sense
of anticipation or stasis, creating foregrounds and backgrounds, and so on. When such signs as ‘record’ or
‘bridge’ are given, what is recorded may not be used, or a bridge might develop into its own section without
ever returning to the original ‘A’ section. Moreover, the use of such signs creates the very possibility of an
expectant ‘return’ of a section after a divergence, thus establishing its own kind of temporality, with
sections that can be repeated instead of a constantly shifting and transforming musical texture.

The students generally stated that one of the main skills that Kobranie had taught them was to develop an C25.P36
awareness of the overall shape of the music that they are making or processing. Many who did not really
improvise before starting the course stated that they had learned to improvise, but to do so in a way that
showed what they called a ‘compositional mindset’. As one student put it:

I think it teaches you to think more abstractly, or on di erent levels, musically. To be more C25.P37
concerned, when you’re improvising, with form and function, and that you have a much bigger
picture of the piece than when you’re improvising without a processor. So it feels, for me, as
though it has given me much more direction in my improvisations, especially with a larger group.
7
(HA)

The form of the piece, however, is not given solely by the processor. The students soon nd out that having C25.P38
a predetermined direction in mind is usually not conducive to making a good piece. Although it can be useful
to have an idea of how to begin, the signs are imprecise in such a way that usually the processor will be
unable to predict the musical outcome:

The ideas of the musicians are very valuable, and you can build on them. When you are conducting C25.P39
you can say in general terms, ‘I would like to hear a rhythm now’, but you cannot specify it in
detail. That is not your job, either; it is more uid when things happen naturally, when things start
8
to emerge. (AD)

p. 538 As these musicians explain, the feeling of a piece of music starting to have a sense of direction is something C25.P40
that emerges from a relational, interactive process. The signs that they employ do not determine the course
of action, but are themselves part of this interactive process. If the ICP’s scores function as maps, Kobranie
only has signposts.

The processor, rather than establishing a ow of duration, has to be very attuned to the possibilities for C25.P41
shaping form inherent in the various contributions already given by the various musicians. When one
student, after a less successful piece, commented: ‘I didn’t know where I was going with the music’, Olthuis
redirected her attention to what was already happening: ‘The music already was somewhere.’ When
commenting favourably on one of the recordings I had made during a class, Olthuis said:

He is actually very modest, he doesn’t do so much. He is turning people o , loosening things up— C25.P42
he’s making sure the water keeps owing. He can see it when there’s something impeding the ow
9
of the river, or where he needs to build a little dam so that the ow can be controlled.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
Hence, the musical ‘ ow’ is not an inner mental state, but emerges from the interactions between the C25.P43
musicians. The Kobranie signs are means to negotiate these interactions and the temporal possibilities
inherent in them, and thus participate in the construction of musical time.

The Genetic Choir C25.S5

While the ICP has a large repertoire of notated pieces, and the Kobranie method consists of a system of hand C25.P44
signals and gestures, the Genetic Choir is more hesitant about committing itself to such a determined
system. Its director, Thomas Johannsen, describes its work as aiming towards a kind of ‘swarm
intelligence’, where the direction of the music is not imposed externally, but results from the equal
participation of di erent voices, a process that gives rise to self-organizing systems and emergent
structures resulting from the interaction between these voices (see Borgo 2005; Sawyer 2003 for similar
ideas about group creativity and improvisation). A frequently used image by Johannsen is that of a ock of
birds moving as one: a formation of movement resulting from the mutual coordination of individual birds
but not reducible to them. Hence, there is something of a resistance to ‘notation’ in a traditional sense
within the choir, as there is in many practices of musical improvisation—in fact, the choir might simply be
described as making freely improvised music, unlike the ICP or Kobranie. However, Johannsen is always
looking to describe or codify certain aspects of these processes, and uses the results from his research as
exercises in workshops or as the basis for projects and performances.

One such workshop exercise is to sing in a duo, with one person directly and exactly imitating the other, C25.P45
then switching these roles around, and then having no ‘leader’ but simultaneously both following and
p. 539 leading each other. The exercise, like the image of the ocking birds, again suggests that the shared
temporal ow is not quite a matter of a collective consciousness, but of the coordination of processes going
on in the material world. This is also true of other exercises; in ‘any tone’ all singers hold a long note,
consciously attending to their own place within the dense, dissonant harmony that results; in ‘any rhythm’
the singers do the same but with rhythm, attempting to maintain a sense of independence while also being
mindful of the complex polyrhythmic structures that emerge. Exercises in phrasing start by the musicians
moving in pairs or small groups from one point in the room to another at di erent tempos, to acquire a
sense of the momentum resulting from moving together in such a way. As Ralph de Rijke, one of the singers,
put it: ‘Focusing on each of these things might be a way of strengthening certain ‘muscles’ […] It’s a
10
question of developing muscle memory and uidity which are useful in performance.’ The idea of such
exercises as training certain muscles suggests that they are not so much intended to develop a way of
thinking or a form of consciousness as they are to cultivate a physical awareness of the relation to one’s
fellow singers and their movements.

Hence, far from a complete rejection of notation, the fascination with self-organizing systems in C25.P46
Johannsen’s case has led to the use of a range of di erent, contingent representations of musical processes.
Some are simple games or instructions, others describe certain musical parameters that can serve as a point
of focus for the singers. During my work with the group, they were involved in a project on the concepts of
Time, Texture, and Meaning. These were considered to be aspects inherent to any sound, and could be used
by the singers as a way of attending to the music they were making and of imagining ways to transform it.
They also functioned as a kind of research object: in the rst rehearsals, the singers improvised pieces
concentrating on each individual concept, and discussed how this had shaped their creative process and the
relation to their fellow musicians. After a piece on ‘time’, in which the singers experimented with
repetition, silence, and the lengthening and shortening of phrases, for instance, they discussed how their
experience of time was related to their sense of movement. After a piece on ‘meaning’, which featured
references to musical idioms as well as real-world sounds, they discussed how they were guided by visual
imagination in their improvisations, or whether such thoughts distracted them. In the course of the project,
the concepts remained objects of individual and collective experimentation, and the musicians were at

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
liberty to develop their own relation to them, with some developing a clear preference for one concept over
others, and some ultimately not really using them in performance at all. In fact, Johannsen himself
acknowledged that ‘the danger of these kinds of systems is that you become too distracted by your own
11
thoughts, and then you are no longer focusing on the music’, adding that he only used them at moments
where he needed to reconnect to the musical situation. Thus, these concepts are attempts to describe or
represent aspects of musical sound, not as a way to monothetically ‘capture’ them, but as a means for
musicians to relate to the emergent development of musical structures.

Martine van Ditzhuyzen, one of the Genetic Choir singers, suggested that the group’s use of these concepts C25.P47
allowed her to negotiate the ow of musical sounds, as well as her own musical imagination, with a greater
sense of autonomy:

p. 540 Suppose you have a river. You have to cross the river. And that river stands for all the possibilities C25.P48
and thoughts that ash through your head while improvising, all the possibilities of possibilities.
And if there are some rocks, you can use them to cross the river. So in our case that’s Time,
Texture, and Meaning. So it helps you to think, I’ll jump to this rock rst, and then this, and so on.
If you can do that with some calm and attentiveness, you can reach the other side with a feeling of
ow. Alternatively, you can say, there are no rocks. I’ll swim. I’ll join the ow of the river, and
that’s how I get to the other side. But then, I might end up all the way over there! With the rocks I
have some choice where I want to go, and how I want to relate to what the others are doing.
12
Otherwise I’ll swim and the current just sweeps me away.

Van Ditzhuyzen’s metaphor implies that being swept away by the ow of a musical stream of consciousness C25.P49
may not always be desirable. The use of the choir’s concepts gives her a clearer sense of the musical
situation and how she can respond to them. Thus, rather than a sense of detachment and individualism, this
use of notation a ords the singers an e ective way of relating to each other’s actions.

Reconsidering Halbwachs C25.S6

As already mentioned, part of Schütz’s argument is a rejection of the work of Halbwachs, who (according to C25.P50
Schütz) identi ed the content of musical communication with the musical structures represented by a
score. From his point about there being no di erence between the forms of musical interaction in a string
quartet and those in a group of jazz musicians, Schütz thus concludes that ‘these examples simply give
additional support to our thesis that the system of musical notation is merely a technical device and
accidental to the social relationship prevailing among the performers’ (Schütz 1951: 96). Cook’s argument
for the study of musical performance as a form of social interaction is premised on a similar rejection of the
idea that the performance of notated music is simply the transmission of the musical patterns already
contained therein. As he puts it: ‘Writing sucks time out of music’ (Cook 2013: 248). As far as the work-
centred conception of music in traditional musicological scholarship is concerned, this is accurate, and this
is essentially Cook’s main point: that the understanding of performance as reproduction is a confusion of a
‘literary’ appreciation of music with a performative one.
However, from the point of view of performers using notation in their creative process, is it accurate to say C25.P51
that notation is ‘merely a technical device’ and is entirely accidental to their social interactions? The case
studies discussed here certainly suggest otherwise, and in classical music, too, we could argue that the close
coordination of timing among members of a string quartet is an outcome of following the instructions given
by their parts—which is quite a di erent thing from saying that they are reproducing what is ‘in’ the score,
but di erent also to the idea that they are ‘tuning into’ a given arrangement of tones in inner time, since
p. 541 they are themselves arranging these tones through their mutual coordination. In the cases described in

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
this chapter, musicians use forms of symbolic representation of musical ideas, not in order to capture these
ideas so that musicians can ‘reproduce’ them, but rather as a means for musicians to construct and
negotiate di erent kinds of temporality in performance. Indeed, the descriptions of processes of
improvisation in all three practices are quite far from a collective ‘ ow’ of sounds arranged in inner time.
Rather than a ‘tuning into’ a current of consciousness, the process is better described as a ‘tuning’, a
13
rendering compatible of the di erent kinds of movement and activity in the course of performance. The
notations are not ‘merely technical devices’ but tools for creating and managing such heterogeneous
temporalities.

Bruno Latour, arguing that time and space are ‘consequences of the ways in which bodies relate to one C25.P52
another’ (Latour 1997: 176, emphasis original), imagines two people travelling the same distance. One has
to cut her way through a dense jungle, while the other takes a smooth train journey. It is the latter, he
argues, that can make the distinction between subjective and objective time, because of an apparent
disparity between transportation and transformation. The smooth alignment of the tracks, the machinery
of the train, the carefully constructed timetables, and the inner workings of the train companies, to the
traveller who is able to ignore these things, create the illusion of transportation without transformation—
and the misunderstanding that the seemingly short period of time is a product of her mind rather than real-
world events (pp. 174–179). Just like Latour’s train traveller, then, Schütz’s experience misconstrues the
temporality of musical experience as a purely inner one because of a misrecognition of the work done by the
performers to transport him musically. Schütz’s musicians are considered rst and foremost as listeners,
and this listening is characterized as an inner experience, even if it is shared between two or more people.
The material world is necessary because co-performers usually have to be in the same room, and it is useful
for them to follow each other’s gestures and facial expressions, but it does not actually participate in the
making of music.

The disagreement between Schütz and Halbwachs does not so much concern music or musical performance, C25.P53
but can be understood more fundamentally as a debate about the merits of the philosophy of Henri Bergson.
Whereas Schütz conceives of music as essentially happening in Bergsonian ‘inner time’ (as interpreted in a
phenomenological framework), much of the work of Halbwachs to which he is responding was formulated
as a critique of this very concept. Halbwachs’s essay on the collective memory of musicians (1968) was a
rst sketch of a more elaborate theory of collective memory. In this work, which is un nished because he
was imprisoned and killed by the Nazis, Halbwachs argues that individual memory is derived from socially
constituted, collectively shared memories. He sees the Bergsonian idea that the only true reality is an inner
ow of duration as a fundamental isolation of the individual, which makes it impossible to see how
memories can be shared, and thus how mutual understanding is at all possible, because any externalization
or representation of this ow can only distort it.

Indeed, the debate between Schütz and Halbwachs is essentially about the question whether music (and by C25.P54
p. 542 implication human sociality) is based on an unmediated intersubjective consciousness or is
fundamentally mediated. Cook’s reading of this debate in terms of the ‘paradigm of reproduction’ is a
valuable argument in the context of the performative turn in music scholarship, but does not quite do
justice to Halbwachs’s concern (and, as we have seen, Schütz himself might be seen as perpetuating this
paradigm to a signi cant extent). Halbwachs does not consider the score a representation of the music. He
makes clear that ‘these signs [on a musician’s part] are not images of sounds, reproductions of the sounds
themselves. Between these points and traces that we see and the sounds that we hear, there is no natural
14
similarity’ (Halbwachs 1968: 172, my translation). Rather, much like Cook’s suggestion to see scores as
‘scripts’ for social interaction (Cook 2001), Halbwachs aims to describe the role of notation as an object of
social interaction, rather than placing it wholly outside the socio-musical processes of performance as
Schütz does.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
We might see Halbwachs’s considerations of music notation as a preliminary account of what he would later C25.P55
call the ‘cadre’ that constitutes individual memory, a socially and materially constructed support or
sca olding of the mind. He turns to the example of music—speci cally orchestral music—because
performers in the classical tradition have to accurately remember a large number of notes, played in
complex gurations:

One is forced to thus remember them as such, integrally. Music is truly the only art that imposes C25.P56
this condition, because it develops entirely in time, and attaches itself to nothing lasting, and, in
order to be recovered, has to be recreated without stopping. It is therefore the clearest example of
the impossibility of retaining a mass of memories with all their nuance and detail and precision,
without an appeal to the resources of collective memory.

15
(Halbwachs 1968: 200–201, my translation)

Thus, decades before philosophical ideas of the ‘extended mind’ (Clark 2011; Clark and Chalmers 1998), C25.P57
post-humanist visions of ‘cyborg theory’ (Haraway 1991), or even McLuhan’s notion of media as
‘extensions’ of the central nervous system (McLuhan 2013), Halbwachs writes that ‘the sheet music plays
16
the role of a material substitute of the [musician’s] brain’ (1968: 173, my translation), and that the
orchestra thus has to be understood as a whole rather than as a collective of individuals, since ‘the
musicians and their parts form a totality, and we have to consider this totality to understand the
17
recollection of memories’ (p. 176, my translation). In fact, in his later work Halbwachs gave quite some
consideration to the role of material objects in the constitution of collective memory and identity, arguing
that ‘the familiar image of our external environment is inseparable from our selves’ and that ‘the objects
18
that surround us […] are like a silent and immobile society’ (1968: 130–132, my translation).

Schütz is certainly right to draw critical attention to the problematic implication that only musicians who C25.P58
read have a truly ‘musical’ understanding of music. Although this point is never explicitly made by
Halbwachs, his argument on the relation of musical notation to what he calls the ‘language of music’ is
muddled, and rests on problematic and unclear assumptions about the distinction between society and
p. 543 nature. Still, if we accept his argument that the score co-constitutes the social interaction between
musicians—and the case studies discussed here imply as much—then we can also see why Halbwachs
insists that the di erence between the musical interaction of people reading from sheet music and that of
people who perform a song they know by heart is categorical rather than one of degree. Rather than
following Halbwachs in his apparent assumption that the notation forms a representation of the music
‘itself’, we might see this representational function in a more constructionist way and say that notation
constructs the very idea of ‘the music itself’ through its representation of it, and this ontological work is
part of how it constructs forms of social interaction among musicians. We can see something of such a
constructionist view, stressing the reciprocity between musical ontology and sociality, when Halbwachs
writes:

Because music detaches sounds from the other things that are given through the senses, we C25.P59
sometimes imagine that it detaches us from the material world. […] The world where music would
transport us would then be the inner world. But let’s look a little closer. A combination of musical
sounds only seems to be detached from any object because it is itself an object. This object, it is
true, only exists for the group of musicians. But whatever ensures the existence of a fact, a being, a
quality, if not the agreement between the members of a society, that is to say those who take an
interest in it? […] Musical language is not an instrument invented after the fact to x and
communicate to musicians what some of them spontaneously imagined. On the contrary, it is this
language that has created music. Without it, there would be no society of musicians, just as
without laws there would be no city, there would be no citizens. Far from isolating ourselves in the
contemplation of our internal states, music takes us outside ourselves.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
19
(Halbwachs 1968: 189–190, my translation and emphasis)

Generalizing Halbwachs’s ideas on musical performance from memory to the temporality of performance C25.P60
more generally, I would suggest that the creativity of performers is better understood as a turning outward,
to the movements and rhythms of their fellow performers, rather than a turning inward to a vaguely de ned
conception of ‘inner time’. Notation is not a representation of ‘external time’, and as such essentially a
corruption or distortion of musical reality; rather, its musicality might be understood in terms of its virtual
alignment of temporalities.

This does mean that we should reconsider Schütz’s argument that there is ‘no di erence’ between a string C25.P61
quartet and a group of jazz musicians. Rather than a di erence between a literate and an oral culture, or
between a ‘true’ music and popular entertainment, however, we should be more sensitive to how di erent
forms of social and technological mediation (of which notation can form an important part) constitute
di erent musical cultures. Ideas of a uniform ow of ‘inner time’ are as inadequate to account for the broad
range of ontologies and socialities across musical practices as are such simple binary distinctions. To be
sure, a full understanding of these ontologies and socialities would also require attending to temporalities
that extend beyond musical performance and experience. These may include the biographies of the
musicians and listeners involved, or the relation of musics, genres, and (sub)cultures to conceptions of
p. 544 history (Born 2015). Schütz’s work on such questions of time, music, and sociality, which extends far
beyond the particular arguments discussed here, will undoubtedly continue to play an important role in
such theorizations. In my view, the shortcomings identi ed here indicate the need to reconsider this work,
especially the arguments concerning the ‘pluridimensionality’ of time, from the perspective of Deleuzian
rather than phenomenological interpretations of Bergson. More fundamentally, however, we should
perhaps be more sceptical of claims about music’s special relationship to time, as these are invariably based
on mind-centred conceptions of music that fail to account for the material and social processes involved in
its production and performance.

Notes
1. See also Gell (1992), Munn (1992), and Bear (2016) for general overviews of the field. As Gell (1992) argues, there are of C25.N1
course actual di erences in conceptions of time between di erent cultures; the issue is rather how to account for them
anthropologically, and so the real problem is methodological rather than ideological.

2. The latter two case studies are part of a larger project forming a comparative study of the role of notation as a source of C25.N2
musical creativity across di erent musical practices. See Schuiling (2019b).

3. Interview with the author, 20 Feb. 2012. C25.N3

4. Interview with the author, 4 Jan. 2013. On the di iculty of making an ending in jazz performance, see Do man (2011); C25.N4
Dueck (2013).

5. Interview with the author, 31 Jan. 2013. C25.N5

6. The Dutch word branie means something like ʻswaggerʼ. C25.N6


7. Interview with the author, 12 Oct. 2017. Considering that these students are relatively young and only just starting their C25.N7
musical careers, I have chosen to refer to them by initials only.

8. Interview with the author, 23 Oct. 2017. C25.N8

9. Interview with the author, 29 June 2018. C25.N9

10. Interview with the author, 8 Apr. 2018. C25.N10

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
11. Interview with the author, 18 Sept. 2017. C25.N11

12. Interview with the author, 28 Aug. 2017. C25.N12

13. See also Pickering (1995) for an application of this idea to scientific experimentation. C25.N13

14. ʻCes signes ne sont pas des images de sons, qui reproduiraient les sons eux-mêmes. Entre ces trait et ces point qui C25.N14
frappent la vue, et des sons qui frappent lʼoreille, il nʼexiste aucun rapport naturel.ʼ

15. ʻForce est donc de la retenir telle quelle, intégralement. La musique est, à vrai dire, le seul art auquel sʼimpose cette C25.N15
condition, parce quʼelle se développe tout entire dans le temps, quʼelle ne se rattache à rien qui demeure, et que, pour la
ressaisir, il faut la recréer sans cesse. Cʼest pourquoi il nʼy a point dʼexemple où lʼon aperçoive plus clairement quʼil nʼest
possible de retenir une masse de souvenirs avec toutes leurs nuances et dans leur detail le plus précis, quʼà condition de
metre en oeuvre toutes les ressources de la mémoire collective.ʼ

16. ʻLa partition joue donc ici exactement le role de substitut matériel du cerveau.ʼ C25.N16

17. ʻ[…] les musiciens et leurs partitions forment un ensemble, et […] il faut envisager tout cet ensemble pour expliquer la C25.N17
conservation des souvenirs.ʼ Alfred Gell in fact comes close to arguing that ʻtime-consciousnessʼ is simply cognition. See
Gell (1992: 229–241).

p. 545 18. ʻ[…] les images habituelles du monde extérieur sont inséparables de notre moi.ʼ ʻDe fait, les forms des objets qui nous C25.N18
entourent […] sont autour de nous comme une societé muette et immobile.ʼ

19. ʻParce que la musique dégage ainsi les sons de toutes les autres données sensible, nous nous figurons quelque fois quʼelle C25.N19
nous détache du monde extérieur. […] Le monde où la musique nous transporterait serait alors le monde intérieur. Mais
regardons-y dʼun peu près. Une combinaison ou une suite de sons musicaux ne nous paraît détachée de tout objet que
parce quʼelle est elle-même un objet. Cet objet nʼexiste, il est vrai, que pour le group des musiciens. Mais quʼest-ce qui
nous garantit jamais lʼexistence dʼun fait, dʼun être, dʼun qualité, si ce nʼest lʼaccord qui sʼétablit à son suejet entre les
membres dʼun société, cʼest-à-dire entre les hommes qui sʼy intéressent? […] Le langage musical nʼest pas un instrument
inventé après coup en vue de fixer et de communiquer aux musiciens ce que tel dʼentre eux a imaginé spontanément. Au
contraire, cʼest ce langage qui a créé la musique. Sans lui il nʼy aurait pas de société de musiciens, il nʼy aurait pas même
de musiciens, de même que sans lois il nʼy aurait pas de cité, il nʼy aurait pas de citoyens. Loin de nous isoler dans la
contemplation de nos états internes la musique nous fait sortir de nous.ʼ Video recordings of the interviews are available
on the accompanying website
References C25.S7

Abbate, C. (2004). Music—drastic or gnostic? Critical Inquiry 30(3): 505–536. https://doi.org/10.1086/421160 C25.P62
Google Scholar WorldCat

Abel, M. (2014). Groove: An aesthetic of measured time. Brill. C25.P63


Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
Adlington, R. (2013). Composing dissent: Avant-garde music in 1960s Amsterdam. Oxford University Press. C25.P64
Google Scholar Google Preview WorldCat COPAC

Bear, L. (2016). Time as technique. Annual Review of Anthropology 45: 487–502. C25.P65
Google Scholar WorldCat

Bergson, H. (2001). Time and free will: An essay on the immediate data of consciousness (trans. F. L. Pogson). Courier. C25.P66
Google Scholar Google Preview WorldCat COPAC

Birth, K. (2012). Objects of time: How things shape temporality. Palgrave Macmillan. C25.P67
Google Scholar Google Preview WorldCat COPAC

Borgo, D. (2005). Sync or swarm: Improvising music in a complex age. Continuum. C25.P68
Google Scholar Google Preview WorldCat COPAC

Born, G. (2005). On musical mediation: Ontology, technology and creativity. Twentieth-Century Music 2(1): 7–36. C25.P69
https://doi.org/10.1017/S147857220500023X
Google Scholar WorldCat

Born, G. (2015). Making time: Temporality, history, and the cultural object. New Literary History 46(3): 361–386. C25.P70
https://doi.org/10.1353/nlh.2015.0025
Google Scholar WorldCat

Bourdieu, P. (1963). The attitude of the Algerian peasant towards time. In J. Pitt-Rivers (ed.), Mediterranean countrymen: Essays in C25.P71
the social anthropology of the Mediterranean, 55–72. Mouton.
Google Scholar Google Preview WorldCat COPAC

Burnard, P. (2012). Musical creativities in practice. Oxford University Press. C25.P72


Google Scholar Google Preview WorldCat COPAC

Clark, A. (2011). Supersizing the mind: Embodiment, action, and cognitive extension. Oxford University Press. C25.P73
Google Scholar Google Preview WorldCat COPAC

Clark, A., and Chalmers, D. (1998). The extended mind. Analysis 58(1): 7–19. C25.P74
Google Scholar WorldCat

Clarke, E., and Do man, M. (eds) (2017). Distributed creativity: Collaboration and improvisation in contemporary music. Oxford C25.P75
University Press.
Google Scholar Google Preview WorldCat COPAC

Clarke, E., Do man, M., and Lim, L. (2013). Distributed creativity and ecological dynamics: A case study of Liza Limʼs ʻTongue of C25.P76
the Invisibleʼ. Music and Letters 94(4): 628–663. https://doi.org/10.1093/ml/gct118
Google Scholar WorldCat

p. 546 Cook, N. (2001). Between process and product: Music and/as performance. Music Theory Online 7(2). C25.P77
http://www.mtosmt.org/issues/mto.01.7.2/mto.01.7.2.cook.html
Google Scholar WorldCat

Cook, N. (2007). Making music together: Or, improvisation and its others. In Music, performance, meaning: Selected essays, 321– C25.P78
341. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Cook, N. (2013). Beyond the score: Music as performance. Oxford University Press. C25.P79
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
Cook, N. (2018). Music as creative practice. Oxford University Press. C25.P80
Google Scholar Google Preview WorldCat COPAC

Csíkszentmihályi, M. (2008). Flow: The psychology of optimal experience. Harper & Row. C25.P81
Google Scholar Google Preview WorldCat COPAC

Do man, M. (2011). Jamminʼ an ending: Creativity, knowledge, and conduct among jazz musicians. Twentieth-Century Music C25.P82
8(2): 203–225.
Google Scholar WorldCat

Doğantan-Dack, M. (2011). In the beginning was gesture: Piano touch and an introduction to a phenomenology of the C25.P83
performing body. In A. Gritten and E. King (eds), New perspectives on music and gesture, 243–266. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Dolan, E. (2013). The orchestral revolution: Haydn and the technologies of timbre. Cambridge University Press. C25.P84
Google Scholar Google Preview WorldCat COPAC

Dueck, B. (2013). Jazz endings, aesthetic discourse, and musical publics. Black Music Research Journal 33(1): 91–115. C25.P85
Google Scholar WorldCat

Geertz, C. (1976). Person, time, and conduct in Bali. In The interpretation of cultures: selected essays, 360–411. Basic Books. C25.P86
Google Scholar Google Preview WorldCat COPAC

Gell, A. (1992). The anthropology of time: Cultural constructions of temporal maps and images. Berg. C25.P87
Google Scholar Google Preview WorldCat COPAC

Halbwachs, M. (1968). La mémoire collective. Presses universitaires de France. C25.P88


Google Scholar Google Preview WorldCat COPAC

Haraway, D. J. (1991). Simians, cyborgs, and women: The reinvention of nature. Routledge. C25.P89
Google Scholar Google Preview WorldCat COPAC

Heidegger, M. (1962). Being and time (trans. J. Macquarrie and E. Robinson). Blackwell. C25.P90
Google Scholar Google Preview WorldCat COPAC

Ingold, T. (2010). The textility of making. Cambridge Journal of Economics 34(1): 91–102. https://doi.org/10.1093/cje/bep042 C25.P91
Google Scholar WorldCat

Ingold, T. (2012). Toward an ecology of materials. Annual Review of Anthropology 41: 427–442. C25.P92
Google Scholar WorldCat

Ingold, T. (2017). On human correspondence. Journal of the Royal Anthropological Institute 23(1): 9–27. C25.P93
https://doi.org/10.1111/1467-9655.12541
Google Scholar WorldCat
Latour, B. (1997). Trains of thoughts: Piaget, formalism and the fi h dimension. Common Knowledge 6(3): 170–191. C25.P94
Google Scholar WorldCat

McLuhan, M. (2013). Understanding media: the extensions of man. http://public.eblib.com/choice/publicfullrecord.aspx? C25.P95


p=1222206
Google Scholar Google Preview WorldCat COPAC

Monson, I. (1996). Saying something: Jazz improvisation and interaction. University of Chicago Press. C25.P96

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Munn, N. D. (1992). The cultural anthropology of time: A critical essay. Annual Review of Anthropology 21: 93–123. C25.P97
Google Scholar WorldCat

Ong, W. (2005). Orality and literacy: The technologizing of the word. Routledge. C25.P98
Google Scholar Google Preview WorldCat COPAC

Payne, E. (2016). Creativity beyond innovation: Musical performance and cra . Musicae Scientiae 20(3): 325–344. C25.P99
https://doi.org/10.1177/1029864916631034
Google Scholar WorldCat

Pickering, A. (1995). The mangle of practice: Time, agency, and science. University of Chicago Press. C25.P100
Google Scholar Google Preview WorldCat COPAC

Prouty, K. (2006). Orality, literacy, and mediating musical experience: Rethinking oral tradition in the learning of jazz C25.P101
improvisation. Popular Music and Society 29(3): 317–334. https://doi.org/10.1080/03007760600670372
Google Scholar WorldCat

Rink, J. (1990). Review of Review of Musical Structure and Performance, by W. Berry. Music Analysis 9(3): 319–339. C25.P102
https://doi.org/10.2307/853982

Sawyer, K. (2003). Group creativity: Music, theater, collaboration. Erlbaum. C25.P103


Google Scholar Google Preview WorldCat COPAC

p. 547 Sawyer, K. (2017). Group genius: The creative power of collaboration. Basic Books. C25.P104
Google Scholar Google Preview WorldCat COPAC

Schuiling, F. (2019a). The Instant Composers Pool and improvisation beyond jazz. Routledge. C25.P105
Google Scholar Google Preview WorldCat COPAC

Schuiling, F. (2019b). Notation cultures: Towards an ethnomusicology of notation. Journal of the Royal Musical Association 14(2): C25.P106
429–458.
Google Scholar WorldCat

Schütz, A. (1951). Making music together: A study in social relationship. Social Research 18(1): 76–97. C25.P107
Google Scholar WorldCat

Sidran, B. (1995). Black talk: How the music of Black America created a radical alternative to the values of the Western literary C25.P108
tradition. Payback Press.
Google Scholar Google Preview WorldCat COPAC

Sterne, J. (2003). The audible past: Cultural origins of sound reproduction. Duke University Press. C25.P109
Google Scholar Google Preview WorldCat COPAC

Stiegler, B. (1998). Technics and time: The fault of Epimetheus. Stanford University Press. C25.P110
Google Scholar Google Preview WorldCat COPAC

Tackley, C. (2010). Jazz recordings as social texts. In A. Bayley (ed.), Recorded music: Performance, culture and technology, 167– C25.P111
186. Cambridge University Press.
Google Scholar Google Preview WorldCat COPAC

Taruskin, R. (1995). Text and act: Essays on music and performance. Oxford University Press. C25.P112
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471445 by National Science & Technology Library user on 26 May 2023
p. 548 Whorf, B. L. (1964). Language, thought, and reality: Selected writings of Benjamin Lee Whorf (ed. J. B. Carroll). MIT Press. C25.P113
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
CHAPTER

26 Musical Time in a Fast World  C26

Samuel Wilson

https://doi.org/10.1093/oxfordhb/9780190947279.013.27 Pages 549–C26.P99


Published: 08 December 2021

Abstract
In this chapter Wilson addresses the relation between musical temporality and dominant conceptions
of time under recent or ‘liquid’ modernity. He argues that the sonic arts (music, sound art, etc.)
variously withdraw from and/or embrace normative time-making—thereby critically calling into
question our assumptions about lived temporality. Wilson engages two examples, both intimately
connected with the city of New York and the year 1983: Morton Feldman’s minimal yet durational
String Quartet No. 2, and Bill Fontana’s Oscillating Steel Grids along the Brooklyn Bridge, the latter of
which involved sounds from this bridge (tra c, the metal strut work, etc.) relayed live and broadcast
in downtown Manhattan. Both works criss-crossed di erent temporalities and lived rhythms that
contrasted with the speed implicit in 1980s hypercapitalism, forming dialogues between musical time
and the cultures of its production.

Keywords: modernity, postmodernity, acceleration, New York, Morton Feldman, Bill Fontana
Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

WHAT Fredric Jameson (1991) calls the ‘cultural logic of late capitalism’ also manifests particular temporal C26.P1
logics—speci c organizations of time, temporal arrangements of the everyday—that mediate experiences
of time as lived. These are extended throughout a ‘24/7’ existence explored more recently by Jonathan Crary
(2013) and others. It is a truism that music is a temporal art form par excellence and, furthermore, that it
makes and marks time di erently from everyday temporality. I would like to suggest that these two
temporalities face one another: musical time informs everyday temporality, and vice versa, but each cannot
be reduced to the other. This chapter focuses on how society’s dominant temporal rationale—as manifested
in day-to-day lived experience—relates to the temporal experience forged or contoured by music in the
context of a late stage of modernity. This is a relation that cannot be assumed; neither can it be assured,
given the dynamism of musical temporalities, which are multi-layered and open to change, determined not
only by compositional choices but also by new contexts of performance and listening.
It is my argument that music and sound art are time-making activities in which sensitivity to, and C26.P2
experimentation with, temporality—in accordance with or deviation from aspects of the normative temporal
logic of society—might be practised or exercised. To put this slightly di erently: musical temporalities can
problematize and bring attention to socio-temporal logics, from which they di er but to which they are
always related. Here I consider how music and sound art adopt and adapt the temporal conditions of late
modernity.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
It should be noted from the very beginning that while cultural and musical temporalities might contradict C26.P3
one another, late modernity itself provides contradictory temporal modalities. This view is taken against a
widespread narrative of modern and postmodern life, described singularly as a quickening of lived
temporalities, especially after 1970s neoliberalism. This ‘acceleration narrative’—and its revolutionary
p. 550 potential—was embraced by many early twentieth-century modernists, for instance the futurists (see
Highmore 2005: 140–145), and is more recently articulated in accelerationist politics that ‘want to
accelerate the process of technological evolution’ and argue that the one should repurpose the ‘material
platform of neoliberalism’ towards a revolutionary post-capitalism (Williams and Srnicek 2014: 356, 355).

The cultural theorist Ben Highmore argues that while speed and acceleration are undoubtedly aspects of C26.P4
modern life, to focus on speed alone is to miss something crucial. Highmore (2005: 154–157) gives the
example of air travel. This speeds up life, practically shrinking geographical space through speedy
transport; for the passenger, however, the embodied experience of ight is characterized by queues, seated
immobility, and tedium. Highmore calls attention to this idea through invoking Henri Lefebvre’s concept of
rhythmanalysis. Lefebvre (2004) understood everyday life to trace multiple—often con icting—forms of
time; rhythmanalysis took account of this by bringing the observer’s attention to the diverse and
interacting rhythms and structures of time, such as the working day and the factory clock, the cycles of the
seasons, dawn and dusk, fatigue and alertness, hunger and satiation. Highmore suggests that
rhythmanalysis might most productively be understood as a critical sensitivity to the contradictions
inherent in the structuring and living of time under modernity. As Highmore (2005: 154) puts it,
rhythmanalysis ‘refuses to accept the rhetoric of acceleration as adequate to the experience of modernity’.
While I am not performing rhythmanalysis here, I similarly do want to apply a methodological sensitivity to
the contradictory temporality of the everyday, and a resistance to characterizing modernity in the linear
terms of acceleration. Indeed, I argue that these contradictions are made visible and made audible through
artistic sonic practices that instantiate, highlight, co-opt, or transform the temporalities that bear their
traces on bodies, history, and everyday sociality.

Perhaps the paradigmatic image of a hard and fast society—a culture of late capitalist speed, under the C26.P5
1
auspices of neoliberalism—is provided by New York City in the 1980s. The stock exchange, and the free-
market forces of Wall Street, typi ed this quality acutely. Technologically speaking, new computers and
telematics allowed for rapid trades and lightning-fast commodities speculations, whilst the city typi ed a
hub in a global nance system, ‘a national turnstile’, through which the forces of global nance ‘move[d]
megamonies […] at blinding speed’ (to borrow a phrase from Arjun Appadurai, 1990: 298, on global nance
in general); sociopolitically, this was the era of Reaganomics, deregulation, and a time when the individual
had to keep up or lose out. This city, this historical moment, are taken here metonymically, associated with
some of the dominant cultural and temporal logics that were emerging more generally in the late twentieth
century. It is the very fact that 1980s New York evokes such an unambiguous image of speed and intensity
that I hope to tease out through the temporal contradictions of this place/time.

Focusing on two works associated with the city and the year 1983, I argue that music and sound art explored C26.P6
varied and paradoxical modalities of time. In order to demonstrate this, I begin by brie y considering a
work by the sound artist Bill Fontana and its temporal implications. The relationship between time and
music is then explored more fully, before I turn towards my second example, Morton Feldman’s String
p. 551 Quartet No. 2, and more substantially develop the critical themes introduced so far. The contradictory
temporality of Feldman’s quartet is then considered, with Fontana’s work re-emerging as a counterpart to
Feldman’s towards the end of the chapter.

Fontanaʼs Oscillating Steel Grids along the Brooklyn Bridge C26.S1

Bill Fontana’s Oscillating Steel Grids along the Brooklyn Bridge is an example of one of Fontana’s ‘sound C26.P7

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
sculptures’. In art-theoretical contexts, this term conventionally describes sculptures that incorporate
sound as a key element (see Grayson 1975). However, Fontana’s approach is more acute:

In uenced by Duchamp’s strategy of the found object, I began to realize that the relocation of an C26.P8
ambient sound source within a new context would alter radically the acoustic meaning of the
ambient sound source. I conceived such relocations in sculptural terms because ambient sounds
are sculptural in the way they belong to a particular place.

(Fontana n.d., c)

Fontana’s sound sculptures are well known for relocating live acoustic sound sources from one context to C26.P9
another. Elsewhere he has characterized this process very simply: ‘I take sound away from its context’
(quoted in Trebay 1983); sounds produced in one place are broadcast in a new location. Fontana will treat a
bridge ‘like a giant Aeolian harp’, as one reviewer put it of a later UK-based work (Tyne Soundings, 2009),
placing microphones on a structure and relaying the output to another, remote environment (Hickling
2009). Fontana himself has written of this interest in:

hearing the simultaneity of sounds in a natural landscape, a city, a structure such as a bridge, a C26.P10
train station, a harbor or a long stretch of beach. What is so compelling is the natural completeness
of the live ows of musical events and patterns. That the live ambient sound constellations present
such seemingly perfect relationships makes this art form actualize an awareness of what is already
present.

(Fontana 2000[1990], emphasis added)

This nal phrase echoes a Cagean attentiveness to one’s sonic environment, of which one becomes aware C26.P11
through its reframing and defamiliarization. Furthermore, Brandon LaBelle (2006: 233) suggests that
Fontana’s process of relocating and decontextualizing sounds ‘[dislodges] us from a given visual referent
and creat[es] a jag in the perceptual hierarchy of the senses’. This is to say that this awareness might stem
from, and foreground, the sonic dimension of a eld of activity (a landscape, a city) in which the audial is
conventionally relegated as incidental.

p. 552 New York’s Brooklyn Bridge became the basis for one of Fontana’s sound sculptures in 1983, in celebration C26.P12
of the 100th anniversary of the bridge. As such, Oscillating Steel Grids was not a permanent xture, but a
temporary arrangement for actualizing awareness of urban sounding. At this time, the bridge’s roadway
2
was formed partially by a steel grid. This grid, to quote Fontana, ‘ “sang” with oscillating tones whenever
cars moved over its surface’ (Fontana n.d., a). In Fontana’s work, which was in place from May until August,
microphones attached below the bridge’s roadway captured this ‘singing’. The sounds received by the
microphones were transmitted via ‘equalized broadcast-quality telephone lines’ to a new location and a
new context, the Austin J. Tobin Plaza (Fontana n.d., c). This was the plaza that was then directly below the
Twin Towers. Sounds originally from the bridge emanated from eight speakers embedded in the façade of
the World Trade Centre’s Tower One. Loudspeakers were placed on the WTC’s observation terrace, which
overlooked the city and the East River. Fontana writes: ‘The Brooklyn Bridge was clearly visible from this
vantage point and [this] was the rst time I had explored in real time the idea of hearing as far as you could
3
see’ (Fontana n.d., b).

Highmore suggests that early twentieth-century modernism involved an experimentation with, and C26.P13
sensitivity to, questions surrounding time. I want to consider how this logic might be extended to artwork of
a late modern or postmodern age. Fontana’s work highlighted quotidian temporality and also (I will
suggest) the historical marking of time. Cyclical rhythms of the urban everyday became musically perceived.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
As Fontana pointed out, times of day when there was less tra c on the bridge meant faster cars, resulting in
higher pitches; by contrast, during rush hour, and periods of slower, denser tra c, di erent sonic qualities
were e ected (Fontana n.d., c). ‘It’s best in the early evening,’ Fontana told a newspaper reporter at the
time, ‘there’s less tra c on the bridge and the cars that are on it go faster. That creates greater variations of
pitch’ (quoted in Trebay 1983).

Alongside the everyday, the piece also clearly related to historical time—it was after all a work realized in C26.P14
honour of the bridge’s centenary. But this connection to history was not merely abstract: the material
dimension of the centenary celebrations and the disciplinary act of historical remembrance were
emphasized. Richard Haw (2008: 193), in his cultural history of the Brooklyn Bridge, writes:

On the anniversary itself [24 May] crowds headed to the East River in unprecendented numbers. C26.P15
Speeches were given, a parade was staged […] In the evening, a massive reworks display
culminated the day’s events.

On the centennial day, the sounds of these reworks, and boats’ celebratory foghorns—sonic traces of the C26.P16
concrete act of celebration—intermingled with other sounds from the contemporary everyday. A eld of
sonic activity emerged from both the everyday temporality and historical time, from the sounds of the city
and their transmutation by Fontana’s apparatus: the listener became sensitive both to sounds from urban
life—normally passed by—and to the sonic traces of an act of remembrance, of a century now passed.

p. 553 Fontana’s sound sculpture involves a dislocation of localities—the transposition of one sound source into C26.P17
the space of another. It is perhaps pertinent that a bridge provided the original location of these sounds, a
place which is by its very nature a zone of transition between di erent localities (here, Brooklyn and Lower
Manhattan). But di erent localities also bring with them di erent characteristic temporalities. This may
recall Henri Lefebvre and Catherine Régulier’s observation in their essay on ‘The rhythmanalysis of
Mediterranean cities’, that, ‘[as a] link between spaces, [in many Mediterranean buildings] the stairway
also provides a link between times’ (in Lefebvre 2004: 97). ‘Di erent times’ here denote di erent rhythmic
and temporal habits interior to a dwelling, and those of the street life outside. The stairway, for Lefebvre
and Regulier, connects these two times.

Fontana’s bridge is a site of a number of connections and contradictions: it connects di erent ‘acoustic C26.P18
territories’, to borrow LaBelle’s (2010) term—di erent spaces in which noise is a meaningful element both
experientially and (potentially) poetically. Each of these territories implies particular sonic and temporal
characteristics. It also connects and blends together the di erent temporal registers of the everyday and the
historical. Furthermore, cultural and historical rhythms are connected and blended with natural ones—
recalling Lefebvre’s suggestion that everyday life is a place of overlapping and often con icting rhythms—
variously natural, social, economic, and bodily in origin. The relayed sound from the bridge, oating above
its new setting, the Plaza, was ‘mistaken sometimes for wind’ by listeners, suggests Fontana (n.d., b). More
literally, these sounds themselves captured the impact of cycles day and night on the sonic qualities of the
bridge, which a ected (as observed) the sounds of tra c at di erent times of day. The Bridge—a rei ed
object of adoration in the centenary celebration—also became a liminal space of transition, between
di erent soundings and temporalities.
The ʻProblem of Timeʼ in Music and Culture C26.S2

The relationship between time in music and time in culture is not self-evident. Benedict Taylor (2016: 49) C26.P19
has argued that the ‘problem of time is […] one of the central issues of the modern age’:

With the rise and acceleration of technological and social change, the decline of former religious C26.P20
certitudes and xed points of reference, comes an ever greater awareness of the lack of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
immutability of everything contained within our world of experience.

As times were changing, the notion of time itself changed, and new forms arose. Capitalism’s drive for C26.P21
p. 554 accumulation, production, and consumption, the meta-narratological trope of acceleration already
noted—and, indeed, modernity as such, with its development of ends-focused instrumental reason—might
all imply a characteristic directedness, integral to normative modern temporalities. One could mention here
Lefebvre and Régulier’s observation that under modernity linear forms of time (principally, the time of the
working day, symbolized by the uniform ticking of the clock) came to predominate over—though never
fully obscure—other forms of time, such as the cycle of sunrise and sunset, and the cyclical rhythms of the
body (see Lefebvre and Régulier, in Lefebvre 2004: 72–84). More recently, Jonathan Crary (2013: 15) has
written of how the development of late capitalist society necessitates the promise of an ‘end of sleep’: a
minimizing of natural cycles such as those of rest and recuperation—times at which neither production nor
consumption is possible—and their displacement by an experientially at and temporally homogeneous
‘24/7’ existence.

Karol Berger has considered these notions of modern directedness and changing social rhythms in musical C26.P22
contexts. I turn to Berger’s account here, as he focuses particularly on a historically and culturally speci c
treatment of temporalities, as constituted in part through European art-musical practices. I acknowledge
this as I move to consider Feldman’s String Quartet No. 2, a work which, while despite opening issues of
temporal multiplicity in a twentieth-century North American context, is in its use of the quartet indebted to
this Europeanist genealogy. Berger argues that a changing treatment of time in music was inherent to the
development of musical modernity: foremost musical modernity, which crucially involved the historical
ascendance of a linear model of time—the directedness is captured in the phrase ‘time’s arrow’ (Berger
2005: 18–19). This musico-temporal logic reiterated the linearity of instrumental reason and teleological
thinking found in society at large; a new, teleological directedness typi ed in the sonata of the late
eighteenth century (as encountered in much of the music of Mozart, Haydn, and Beethoven). These sonatas
reiterate a teleological logic in their drive towards dramatic closure, in the overcoming of the dramatic
struggle. The musical performance of linearity thus re ected a newly disciplined linearity of time in culture
at large. This treatment of time is, for Berger, a crucial aspect of the advent of musical modernity. I do not
dispute Berger’s claims about the mediation of musical time by sociohistorical forces (though it should be
4
added that cyclical forms of time in music persisted, coming to mean something new). I ask, instead, a
supplementary question: what happens next? What does the driven, accelerated—though contradictory—
temporal logic of a late twentieth-century context mean for then-contemporary musical explorations of
time?
Feldmanʼs String Quartet No. 2 C26.S3

One answer is provided by the music of Morton Feldman, a composer associated with the New York School C26.P23
of artists and musicians. His late music in particular—many works of which are of great length—provides
durational meditations on the problem of time. His String Quartet No. 2 was written for the Kronos Quartet
p. 555 in 1983 and premiered in Toronto in December of that year. Like much of his late music, this work is a

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
very long, generally slow and quiet piece. In fact, it can take ve-and-a-half or even six hours to play. For
practical purposes—initially, due to the need for a speci c time-slot set aside for a live radio broadcast—
early performances used a shortened (four-hour) version of the score. The quartet was performed by Kronos
six times during the 1980s (Cowperthwaite 1996; Villars n.d.). An even shorter version was produced for the
European premiere at the 1984 Darmstadt Summer Course for New Music (Feldman 2006: 185).

The German music critic and theorist Heinz-Klaus Metzger, in discussion with the composer, told Feldman: C26.P24
C26.P25
‘there is a contradiction between your music and the world in which we live. The world is much louder’
5
(Metzger, cited in Feldman et al. 1972). I would like to suggest that something similar can be said of the
temporal contradiction between Feldman’s music and the world contemporary with it. Indeed, Feldman was
himself interested in the problem of time, both ‘in itself’ and in its relation to listeners. Speci cally, he was
interested in not controlling time, which in his view some composers wished to ‘handle and even parcel out’
(he cites Stockhausen as example); ‘I am not a clockmaker,’ Feldman writes. Instead, he expressed an
interest

in getting to Time in its unstructured existence. That is, I am interested in how this wild beast lives C26.P26
in the jungle—not in the zoo. I am interested in how Time exists before we put our paws on it—our
minds, our imaginations, into it.

(Feldman 2000: 87)

Despite a purported interest in looking beyond human time, Feldman commented elsewhere on his interest C26.P27
in music’s relation to society, and a society of listeners. Asked about the length of the second quartet,
Feldman (cited in Gottschalk 2016: 135) stated: ‘I think that the piece is so long because our attention span is
so short! Five minutes is too long for most people—it’s a serious problem.’ Feldman seems to respond here
to what he saw as the then-contemporary conditions of perception, which I would suggest manifested the
dominant temporal logic: transitory times promote transitory attention. The work’s urgency derives not
from echoing the urgent tempo of the cultural and temporal conditions around it, but instead from bending
6
a speculative and contemplative ear that stands in a di erential relation with these conditions.

Form, Dialectics, and ʻUnfixingʼ C26.S4

The String Quartet No. 2 does not explore the driven directedness that one encounters in Berger’s account of C26.P28
an earlier stage of musical modernity. The parts do not, for instance, synthesize into a linearly developing
‘whole’. Feldman’s own distinction between composition and assemblage is enlightening on this point. He
describes the di erent principles between:

p. 556 constructing a ‘composition’ and that of assemblage, which is more what this quartet is about. A C26.P29
‘composition’ for me forms sentence structures within a beginning, middle, and end […] With
assemblage there is no continuity of tting the parts together as words in a sentence or paragraph.

(Feldman 2000: 196)

In the quartet one hears many elements of material repeated with minor alterations. Larger, static ‘ elds’ of C26.P30
material result from the assemblages of these elements, and these elds appear beside one another—
assemblages in a larger sense—as static states without a beginning, middle, or end. Furthermore, these
fragments do not that resolve in response to one another (nor is it implied that they should need resolving),
and they partake of no commonly underlying musical grammar nor any unproblematic structuring of
musical time (e.g. a conventional syntagmatic arrangement of beginning, middle, and end paradigms).

This is in part achieved through a peculiar treatment of musical time, of—put simply—the (expected, C26.P31
established, apparent) relation of musical moments to one another. Feldman detaches gestures and

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
elements of musical material from an unfolding temporal process. He ‘un xes’ these elements from
established patterns of composition and from normative musical-temporal logic. As Feldman put it in 1965:

[O]nly by ‘un xing’ the elements traditionally used to construct a piece of music could the sounds C26.P32
exist in themselves—not as symbols, or memories which were memories of other music to begin
with.

(Feldman 2000: 35)

Feldman’s un xing also speaks to the work’s dialectical engagement with the dominant temporal C26.P33
conditions of the time: his music un xes musical elements from their standardized arrangement in time—
in microcosm, defamiliarizing normative temporal discipline. As he has stated elsewhere in reference to his
music in general (in 1972, prior to the second quartet): ‘The music seems to oat, doesn’t seem to go in any
direction, one doesn’t know how it’s made, there doesn’t seem to be any type of dialectic, going alongside
it, explaining it’ (cited in Feldman et al. 1972).

Needless to say, Feldman’s temporal experimentation di ers radically from time-making in music of the C26.P34
earlier twentieth and, indeed, the nineteenth century. And this has implications for musical expressions of a
distinctly late twentieth-century subjectivity. As discussed, Berger, considering historical changes in
musical temporality, illustrates that a teleological drive comes to dominate presumptions about the role of
time in music (‘time’s arrow’). It should be added here that, crucially, this linear temporality veiled music’s
embodiment of then-contemporary crypto-dramas of overcoming and subjective self-determination.
Beethoven’s ‘Heroic’ symphonies provide paradigmatic examples of this function (e.g. No. 3 in E♭ major Op.
55, 1802–4, and No. 5 in C minor Op. 67, 1804–8): through the temporal duration of these works, they
deliver an unfolding process connoting the struggle and ultimate mastery over the interior world of the
bourgeois post-Enlightenment subject (Burnham 1995; Schmalfeldt 2011). That linear time and a
p. 557 teleologically determined subjectivity were interlinked should come as no surprise: it is a connection that
is deeply rooted in the classical alignment of time with interiority and space with exteriority. As Fredric
Jameson (2003: 697) puts it, in this view ‘time governs the realm of interiority, in which both subjectivity
and logic, the private and the epistemological, self-consciousness and desire, are to be found.’ Coextensive
7
dialectics, of both music and the self, here unfold temporally.

Surfaces, Moments, and the ʻEnd of Temporalityʼ C26.S5

Given this association between temporality and subjective interiority, Feldman’s rejection of conventional C26.P35
temporality suggests that the quartet disinvites a hearing of subjective interiority as expressed temporally.
8
Instead, in the sense of exteriority, the work also becomes a space from which one listens ‘outwards’.
Furthermore, Jameson has diagnosed late capitalism’s treatment of time as a reduction of experience to
bodily immediacy in the moment, an ‘End of Temporality’:

This situation has been characterized as a dramatic and alarming shrinkage of existential time and C26.P36
the reduction to a present that hardly quali es as such any longer, given the virtual e acement of
that past and future that can alone de ne a present in the rst place.
(Jameson 2003: 708)

Feldman’s quartet, like much of his music, expands moments of time outwards through varied repetitions C26.P37
of musical elements un xed from a linearly directed temporal logic. His musical temporality is situated
within the experiential-temporal conditions diagnosed by Jameson; however, I argue that his music does
not merely reproduce Jameson’s postmodern ‘reduction to a present’ unproblematically.

In addition to his technique of un xing, what allows Feldman to achieve this is a mimesis of the surface. By C26.P38

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
‘surface’, I mean traces of musical material which have become abstracted forms, which resist a simple
reading in terms of symbolic or historical depth, or which act a rmatively as markers of a xed structure
beneath them. This surface evokes a depthlessness famously suggested by Jameson and others as
characteristic of the postmodern. At the same time, this surface conjures something akin to the ‘ atness’
9
Clement Greenberg (2003) famously identi ed in modernist painting. (Although, unlike Greenberg’s
aesthetics, this is committed to an experience beyond a neo-Kantian appreciation of form—as explored
below, this has implications for things like the body.) Feldman’s mimesis is constituted by a tracing of
elements of gestures and materials that recall but never fully embody their historical associations—for
example, could-almost-be-tonal pitch collections and fragments of repeating motivic material. And, as
already noted, Feldman un xes these mimetic traces from their established patterns of composition and
hearing.

Mimesis of the surface comes through in the ‘degrees of stasis’ explored; un xed elements are given C26.P39
time/space in which to speak (although, given their un xing from codi ed ways of hearing, one cannot be
p. 558 sure what they are saying). In his essay on ‘Crippled symmetry’, dated two years before his quartet,
Feldman noted that he discovered this idea of stasis—the temporal kin of space’s ‘ atness’—through
painting:

Stasis, as it is utilized in painting, is not traditionally part of the apparatus of music. Music can C26.P40
achieve aspects of immobility, or the illusion of it: the Magritte-like world Satie evokes, or the
‘ oating sculpture’ of Varèse. The degrees of stasis, found in Rothko or a Guston, were perhaps the
most signi cant elements that I brought to my music from painting.

10
(Feldman 2000: 149)

Feldman (quoted in Gottschalk 2016: 137) stated something similar of his Frank O’Hara (1973): ‘My primary C26.P41
concern (as in all my music) is to sustain a ‘ at surface’ with a minimum of contrast.’ Sustaining a ‘ at
surface’ and an intense focus on un xed musical elements somewhat evacuates Feldman’s musical
materials of their conventional signi cance, both in terms of paradigms (and their deviations) and in the
ways they (‘should’) t together syntagmatically. As a result, unlike the processual unfolding of works that
articulate themselves temporally around some (albeit illusive) teleological objective, a desired goal is
evoked neither through directed musical motion nor through a dialectical struggle for a rmation. This
brings an intense, almost meditative focus on the moment, chiming with Heinz Knobeloch’s
characterization of another of Feldman’s late works, Neither (1977); he suggested that instead of hearing an
unfolding, large-scale form, we experience something more akin to the turning of the pages of a diary
(Knobeloch, quoted in Johnson 2006: 77).
Figure 26.1

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
Morton Feldman: String Quartet No. 2 (1983), p. 60, beginning of the second system. ©Copyright 1983 by Universal Edition C26.F1
(London) Ltd., London/UE17650.

Reproduced with permission.

Figure 26.1 provides an excellent example: a passage in which an un xing of elements and focus on the C26.P42
surface is explored in explicitly temporal terms. The same time signatures appear across all four
instruments simultaneously, as they do conventionally. However, their placement deviates from the norm.
Within this static eld, each instrument takes its own time signature; although, given that each takes
varying combinations of the same four time signatures, on completing the eld they reconvene into a
p. 559 mutually held common time (a silent two-two bar). Note that in the score, despite the bars being of
di ering lengths, the spacing of the bar lines (and the notes within them) on the page do not re ect the
sonic result produced, in which the players do not begin and end their bars at the same time. A sense of
temporal depth in the score is lost as each bar, in each instrumental part, signi es a framing of time that is
removed from the temporal ow at large. The time signatures do not frame the temporal ow but rather
exist in the space of this larger assemblage, this static eld. Furthermore, in a painterly sense, the score
itself becomes a surface written on, a surface on which temporal elements are un xed and then distributed,
11
rather than developed so as to unfold temporally.

Feldman’s music embraces duration through minute variations and repetition. He himself has contrasted C26.P43
the concepts of variation and repetition (reiteration), and considers his music as exploring a synthesis of
the two (see Gottschalk 2016: 143). These have a direct impact on the listener’s perception, on how one
hears any moment as relating to another. Feldman commented directly on this perceptual implication in
relation to the compositional process of the quartet:

In my [second] string quartet I often do things to alienate memory. For example, I might have C26.P44
something return, but it returns in a di erent ordering. It seems only a little familiar. Like when
you see someone for the rst time after ve years and she looks like the same person but […] So I
put things into a di erent ordering. Some material might even return in another key, god forbid,
which is evoking the whole idea of modulation.

(Feldman 2006: 186; see also Gottschalk 2016: 144)

As is evident from Feldman’s allusion to modulation, there is a tension in his accounts of un xing and C26.P45
12
variation/repetition. On the one hand, un xing suggests as a liberation of surface materials from their
tethering to normative temporal logic: they exist to be heard for themselves, on their own terms. On the
other hand, the compositional organization (or assemblage) of these same materials seem to echo pre-
established musical processes, resonating associations, for instance, of (tonal) harmonic practices (in a
post-tonal context). Put another way, un xed materials seem to be foregrounded as existing self-evidently
for themselves; at the same time, their temporary—not xed—organization in work, evokes traces of older
musical practices of organizing music’s elements in time (the logic of exposition and return, modulation,
and so on).

One could re ect on how these dialectics of surface material and temporal logic, transitory presentness and C26.P46
past associations, relate to the normative temporal conditions of the late twentieth century. I suggest that
Feldman’s focus on un xed moments is not a strategy of a postmodern reduction of experience to the
moment, like that diagnosed by Jameson in contemporary society in general. It is, instead, a lengthy

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
meditation on the postmodern reduction of experience to the moment. The kind of experiential context
found in the ‘throwaway’ society of late capitalism—where the temporariness of consumer products seems
itself to become an atemporal truth; the way things are, will be, and should be—is symbolically resisted in
the creation of a sonic time/space antithetical to the temporal logic here implicit. This is a time/space that
p. 560 eschews the dominance of administered experience through an un xing of its materials from their
standardized arrangement, one in which the depthless surface becomes a critical resource for bringing to
into focus the very cultural logic that materials’ arrangement they normatively manifests. Zygmunt
Bauman (2012) suggests that crucial to this period of recent modernity—what he calls ‘liquid modernity’—
is a feeling and embrace of transitoriness (also see Wilson 2021). Feldman’s music engages this situation
dialectically—his sustained musical meditation negates the transitory—yet, without committing to
anything solid and unchanging, the music at the same time embraces it. A paradoxical sustained transience
results. ‘Marking and measuring the dimensions of its own vanishing, insistently demarcating a
disappearance, the music might nally, ephemerally, be heard as the sedimented sound of time, time itself
sounding,’ as Clark Lunberry (2006: 25) put it. As such, Feldman’s quartet provides a di erent timeframe—
a di erent framing of time—from those of the prevailing temporal conditions of that of the hard-and-fast
capitalism of 1980s New York, a time and place that metonymically captures broader social, aesthetic, and
temporal modern trajectories.

Contrasts and Conclusions C26.S6

Sonic Heterochronies C26.S7

In a characterization that has implications for music and sound art beyond merely Feldman’s and Fontana’s C26.P47
1983 works, one could productively frame these examples in relation to terms o ered by Henri Lefebvre and
Michel Foucault. These thinkers each re ected on time/space and its relation under forms of sociality
contingent on the historical and material development of modernity.

Speci cally, Lefebvre and Régulier recognize a temporal category that they label ‘appropriated’ time, a time C26.P48
created or gifted, not a time that one is obliged to follow through its imposition:

Whether normal or exceptional, it is a time that forgets time, during which time no longer counts C26.P49
(and is no longer counted). It arrives or emerges when an activity brings plenitude, whether this
activity be banal (an occupation, a piece of work), subtle (meditation, contemplation),
spontaneous (a child’s game, or even one for adults) or sophisticated.
13
(Lefebvre 2004: 76)

Importantly, however, Lefebvre and Régulier suggest that this is a time that ‘is in time: it is a time, but does C26.P50
not re ect on it’ (Lefebvre 2004: 77, emphasis original). This is where critical artistic practices might enter:
what these share with Lefebvre’s appropriated time is that, for the time that they are ongoing, the normal
rules of time seem not to apply. However, in contrast with Lefebvre’s appropriated time, music can thereby
p. 561 become a relatively autonomous time/space of re ection on time’s appropriation for purposes outside
the normative. Rather than appropriated time—a time outside time—which is ‘in harmony with itself and
with the world’ (Lefebvre and Régulier, in Lefebvre, 2004: 77), dialectical appropriations of time, such as
Feldman’s, might both trace e ects of normative time (perception, attention, memory) while
simultaneously asserting a time sometimes harmonious and sometimes dissonant to dominant temporal
logics.

Complementing this view of ‘another’ time, one might also consider the temporal dimension of what C26.P51

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
Foucault calls ‘heterotopias’, spaces ‘other to’—held in relation to—the norms of society: these
heterotopias either ‘create a space of illusion that exposes every real space’ or ‘create a space that is other,
another real space, as perfect, as meticulous, as well arranged as ours is messy, ill constructed, and
jumbled’ (Foucault 1986: 27). As Foucault (p. 26) also notes, heterotopias ‘are most often linked to slices in
time—which is to say that they open onto what might be termed, for the sake of symmetry, heterochronies’.
In Foucault’s view, some such spaces, for example museums and libraries, accumulate time, gathering pasts
into a present moment of understanding. In contrast, others posit ‘time in its most eeting, transitory,
precarious aspect’, such as vacation spaces and mobile fairgrounds (p. 26).

Feldman’s and Fontana’s works seem to hold converse relations under Foucault’s guration of the C26.P52
heterotopia/heterochronia. As already argued, Feldman’s is eeting (present tense), though in a manner
that demands attention to this very condition of eetingness, defamiliarizing musical time. Paradoxically,
the String Quartet No. 2 is, at the same time, repeatable and re-performable. Fontana’s work was (past
tense) quite the opposite: it drew attention to the daily rhythms of the city and historical memory, in a
manner that naturalized and celebrated these things. But, unlike Feldman’s quartet, as a singular temporary
installation attached to a speci c historical moment, the work is itself unrepeatable.

It should also be remembered here, in light of the Foucauldian heterotopia, that holding a relation to the C26.P53
normative is not to escape normative time-making in an absolute sense. Crucially, times and spaces ‘other
to’ the dominant might, potentially, only supplement and strengthen it. This is clearest when the
(temporary) freedom they o er is instrumentalized as belonging to a time/space of recuperation, rest, or
carnivalesque ecstasy, from which one might later emerge refreshed to ful l one’s normative duties with
renewed energy. The classical music tradition, as Ian Biddle has noted, is a relationally constituted category,
a eld measured in its di erence to other sets of musical (and implicitly sociopolitical) practices. It has
‘always been (and imagined as being) a kind of island or nature reserve in which delicately sanctioned
musical practices and rituals must be preserved, held in place, against the thump and roar of a brash
modernity’ (Biddle 2011: 3).

Employing a quintessentially classical medium in the string quartet, Feldman’s music bears traces of this C26.P54
lineage. His music provides a sanctuary from the temporal logic of late capitalism. While this is true, at the
same time this other space is anything but unproblematic. His island bears modernity’s temporal debris on
its shores; questions of attention, memory, and the transitoriness of liquid modernity are explored
p. 562 sonically, through (for example) a mimesis of the surface and an un xing of musical materials.
Furthermore, one is always unsure of the stability of this island underfoot: its materials are fragments
constituting assemblages rather than a wholeness arrived at teleologically—denying listeners the illusion of
ever treading solid ground. This perhaps recalls Daniel Chua’s reading of Adorno’s dialectics as ‘drifting’:
that the immanence of Adorno’s critique highlights that ‘there is no vantage point from which to
philosophize’, and that ‘to drift is to loose all bearings’ (Chua 2006: 2). In similarly eschewing a unitary
whole, con rmed by a culminating moment of desired arrival, Feldman grants no privileged position or
vantage point at which perception might become retrospectively concretized and solidly grasped.
Performance/City, Times/Spaces C26.S8

Feldman’s and Fontana’s di erent times also manifested through di erent audience and performer C26.P55
experiences in each case, experiences themselves related to the temporal and material conditions of then-
contemporary modernity. Further, this experiential dimension also points to the body’s place within
performance/city times/spaces. In accordance with the concert tradition, in Feldman’s work an audience
gathers together in an act akin to ritual. The performers are physically present; their bodies are present—

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
and bodies tire over the hours of performance. Indeed, the Kronos Quartet had to cancel their performance
of the quartet at the Lincoln Center Festival 96. A press release (Cowperthwaite 1996) states that in
Feldman’s work ‘the musicians never rest or put their instruments down: they play continuously for the
duration of the piece. In preparation for this performance, members of the Quartet have experienced serious
physical side e ects as a result of the unusual nature of this work’ (also see Lunberry 2006). The music
requires repetitive musical gestures constituting intense physicality and extremes of bodily self-discipline.

Fontana’s work, by contrast, does not gather together an audience or performers as such. It does not draw C26.P56
individuals into a delineated performance space. Sound is heard within public spaces, the Plaza and the WTC
observation deck, yet not contained within these spaces. As noted already, the problematic of locality is
crucial to the conception of Fontana’s project. Individual listeners do not constitute an audience proper,
contrasting with Feldman’s music, in which he gives us a quartet of performers towards which an audience
collectively directs its attention. The dispersal of individual listeners in Fontana’s work is perhaps true to
the spirit of the modern city, and in particular the experience of a speci c form of urban space as found in
the case of the Plaza. Bauman refers to spaces such as these as ‘public yet non-civil space[s]’ (Bauman 2012:
97); individuals pass one another—yet they do not engage one another—in a space that is nonetheless
public. Lefebvre and Régulier (Lefebvre 2004: 75) similarly emphasize the temporal in this modern yet
public solitude: ‘In one day in the modern world, everybody does more or less the same thing at more or less
the same times, but each person is really alone in doing it.’ Each experiences the temporal complexities of
city life, simultaneously though separately. Further, whereas human physicality—and embodiment of
p. 563 repetitive gestures—is crucial to the signi cance of temporality in Feldman’s quartet, it is the forces of
both human and non-human materialities (see also Wilson 2017; 2018) that produce the sonic outcomes in
Fontana’s piece: the transport of commuters and commodities across the Brooklyn Bridge, and their
uctuation and rhythming dependent on the daily repetition of business hours, commuter tra c, and
natural cycles of day and night.

To make time in music is to manifest contested notions of temporality—notions related to but distinct from C26.P57
the society in which this music is made. Music renders sensible time and its contradictions. To borrow the
words of Jacques Attali (1985: 4), ‘Music is more than an object of study: it is a way of perceiving the world.’
And, as a means of perception, I suggest that it is both a record of and reaction to historically changing
temporal conditions.

Acknowledgements C26.S9

This chapter has had a long gestation, developing aspects of material on Feldman’s String Quartet No. 2 C26.P58
originally appearing in my PhD project. I am very grateful to Chrysi Papaïoannou for her generous feedback
on aspects of this material, and to Toby Young and an anonymous reviewer for their thoughtful comments
on the chapter as a whole.
Notes

1. Marie Thompson has suggested that ʻa distinction between ʻpre-ʼ and ʻpost-ʼ neoliberal music cultures is, in at least some C26.N1
geographical locations,ʼ an important historiographical distinction. The image of NYC is thus taken here as manifesting
one moment in this cultureʼs emergence. See Thompsonʼs contribution to David Clarke (2018: 411–462), at 455.

2. In the late 1980s this was paved over. C26.N2

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
3. Fontanaʼs webpage on his work (Fontana n.d., b) includes an audio excerpt. Post-9/11, the work has come to mean C26.N3
something retrospectively di erent in its relation to the WTC and historical time.

4. One should note that linearity is not the whole story of the historical emergence of modalities of musical time. While linear C26.N4
directness concretized in late 18th- and early 19th-c. music, many cyclical treatments of time arose in the later 19th c.
(Schumann, Mahler, etc.). Given my focus on music in the context of the temporal logic of the late 20th c., narratives of
acceleration, and driven capitalist speed, I foreground directedness in this discussion. On cyclical time and musical form,
see ch. 5, ʻLa sonate cyclique and the structures of timeʼ, in Taylor (2016).

5. The transcript of Feldman, Brown, and Metzgerʼs discussion was taken from a recording released in the 4-LP set ʻMusic C26.N5
before revolutionʼ (EMI Electrola, 1C 16528954/957, 1972). Christian Wol (cited in Gottschalk 2016: 24) has similarly stated
of quiet music in general: ʻTo be with this music is to find a kind of refuge from the violence of the times. But then the real
strength of quiet music would be to make that refuge a waystation (there are no refuges): to begin to undo and unmask
that violence.ʼ

p. 564 6. One could argue that awareness of these conditions became only more urgent a er Feldmanʼs time, in which attention C26.N6
itself has become subject to new forms of economic rationality: ʻIn the late 1990s, when Google was barely a one-year old
privately-held company […future CEO] Dr. Eric Schmidt declared that the twenty-first century would be synonymous with
what he called the “attention economy,” and that the dominant global corporations would be those that succeed in
maximizing the number of “eyeballs” they could consistently engage and controlʼ (Crary 2013: 75).

7. Daniel Chua (2011) has further considered this issue of space and time, and objective exteriority and subjective interiority C26.N7
in Beethovenʼs music.

8. As Jameson (2003: 697) points out, it is never simply the case of the one strategy or the other (either time-interiority or C26.N8
space-exteriority): ʻsuch descriptions are clearly predicated on the operative dualism, the alleged historical existence, of
the two alternatives.ʼ

9. I am reading Greenbergʼs (2003) idea against the grain; Greenberg identified this characteristic flatness, or at least its C26.N9
possible limit, as unique to painting.

10. Furthermore, Catherine Laws (2009) has explored connections between Feldmanʼs music and the paintings of Jasper C26.N10
Johns.

11. This notion of score as space of distribution recalls Theodor Adornoʼs (1995: 67–68) suggestion that in some music, ʻTime C26.N11
[…] is planned, disposed of, organized from the top down as a whole, as only visual surfaces once where.ʼ

12. Indeed, as is probably apparent already, Feldman made numerous contradictory statements about his music. C26.N12

13. In fact, music o en fulfils description of appropriated time: ʻa time that forgets time, during which time no longer counts C26.N13
(and is no longer counted)ʼ (Lefebvre and Régulier, cited in Lefebvre 2004: 76).
References C26.S10

Adorno, T. (1995). On some relationships between music and painting (trans. S. Gillespie). Musical Quarterly 79(1): 66–79. C26.P59
Google Scholar WorldCat

Appadurai, A. (1990). Disjuncture and di erence in the global cultural economy. Theory Culture Society 7: 295–310. C26.P60
Google Scholar WorldCat

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
Attali, J. (1985). Noise: The political economy of music (trans. B. Massumi). University of Minnesota Press. C26.P61
Google Scholar Google Preview WorldCat COPAC

Bauman, Z. (2012). Liquid modernity. Polity Press. C26.P62


Google Scholar Google Preview WorldCat COPAC

Berger, K. (2005). Timeʼs arrow and the advent of musical modernity. In K. Berger and A. Newcomb (eds), Music and the aesthetics C26.P63
of modernity, 3–22. Harvard University Press.
Google Scholar Google Preview WorldCat COPAC

Biddle, I. (2011). Music, masculinity, and the claims of history: The Austro-German tradition from Hegel to Freud. Ashgate. C26.P64
Google Scholar Google Preview WorldCat COPAC

Burnham, S. (1995). Beethoven hero. Princeton University Press. C26.P65


Google Scholar Google Preview WorldCat COPAC

Chua, D. (2006). Dri ing: The dialectics of Adornoʼs Philosophy of new music. In B. Hoeckner (ed.), Apparitions: New perspectives C26.P66
on Adorno and twentieth-century music, 1–17. Routledge.

Chua, D. (2011). Listening to the self: The Shawshank Redemption and the technology of music. 19th-Century Music 34(3): 341– C26.P67
355.

Clarke, D. (2018). Defining twentieth- and twenty-first-century music. Twentieth-Century Music 14(3): 411–462. C26.P68
Google Scholar WorldCat

p. 565 Cowperthwaite, J. (1996). Kronos Quartet cancels performance of Morton Feldmanʼs String Quartet II. Press release 29 July 1996. C26.P69
http://web.archive.org/web/20030824095235/http://arts.endow.gov:80/artforms/Music/KronosPR.html
WorldCat

Crary, J. (2013). 24/7: Late capitalism and the end of sleep. Verso. C26.P70
Google Scholar Google Preview WorldCat COPAC

Feldman, M. (2000). Give my regards to Eighth Street: Collected writings of Morton Feldman. Exact Change. C26.P71
Google Scholar Google Preview WorldCat COPAC

Feldman, M. (2006). Morton Feldman says: Selected interviews and lectures 1964–1987 (ed. C. Villars). Hyphen. C26.P72
Google Scholar Google Preview WorldCat COPAC

Feldman, M., Brown, E., and Metzger, H.-K. (1972). Morton Feldman, Earle Brown, Heinz-Klaus Metzger in discussion, 1972. C26.P73
http://www.cnvill.net/mfmetzgr.htm
Google Scholar Google Preview WorldCat COPAC

Fontana, B. (n.d., a). Borrowed landscapes: The sound sculptures of Bill Fontana 1973–1996. C26.P74
http://www.resoundings.org/PDF/fontanasoundsculpture.pdf
Google Scholar Google Preview WorldCat COPAC
Fontana, B. (n.d., b). Oscillating steel grids along the Brooklyn Bridge. http://resoundings.org/Pages/Oscillating.html C26.P75
Google Scholar Google Preview WorldCat COPAC

Fontana, B. (n.d., c). The relocation of ambient sound: Urban sound sculpture. C26.P76
http://www.resoundings.org/Pages/Urban%20Sound%20Sculpture.html
Google Scholar Google Preview WorldCat COPAC

Fontana, B. (2000[1990]). The environment as musical resource. http://www.resoundings.org/Pages/musical%20resource.html C26.P77

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Foucault, M. (1986). Of other spaces (trans. J. Miskowiec). Diacritics 16(1): 22–27. C26.P78
Google Scholar WorldCat

Grayson, J. (ed.) (1975). Sound sculpture: A collection of essays by artists surveying techniques, applications and future directions C26.P79
of sound sculpture. A.R.C. Publications.
Google Scholar Google Preview WorldCat COPAC

Greenberg, C. (2003). Modernist painting: 1960–65. In C. Harrison and P. Wood (eds), Art in theory 1900–2000: An anthology of C26.P80
changing ideas, 754–759. Blackwell.
Google Scholar Google Preview WorldCat COPAC

Gottschalk, J. (2016). Experimental music since 1970. Bloomsbury. C26.P81


Google Scholar Google Preview WorldCat COPAC

Haw, R. (2008). The Brooklyn Bridge: A cultural history. Rutgers University Press. C26.P82
Google Scholar Google Preview WorldCat COPAC

Hickling, A. (2009). Bill Fontana. Guardian, 24 July. http://www.theguardian.com/music/2009/jul/24/bill-fontana-sage-baltic-tyne C26.P83

Highmore, B. (2005). Cityscapes: Cultural readings in the material and symbolic city. Palgrave Macmillan. C26.P84
Google Scholar Google Preview WorldCat COPAC

Jameson, F. (1991). Postmodernism: Or the cultural logic of late capitalism. Verso. C26.P85
Google Scholar Google Preview WorldCat COPAC

Jameson, F. (2003). The end of temporality. Critical Inquiry 29(4): 695–718. C26.P86
Google Scholar WorldCat

Johnson, J. (2006). Elliptical geometry of Utopia: New music since Adorno. In B. Hoeckner (ed.), Apparitions: New perspectives on C26.P87
Adorno and twentieth-century music, 69–84. Routledge.

LaBelle, B. (2006). Background noise: Perspectives on sound art. Continuum. C26.P88


Google Scholar Google Preview WorldCat COPAC

LaBelle, B. (2010). Acoustic territories: Sound culture and everyday life. Continuum. C26.P89
Google Scholar Google Preview WorldCat COPAC

Laws, C. (2009). Feldman–Beckett–Johns: Patterning, memory and subjectivity. In B. Heile (ed.), The modernist legacy: Essays on C26.P90
new music, 135–158. Ashgate.
Google Scholar Google Preview WorldCat COPAC

Lefebvre, H. (2004). Rhythmanalysis: Space, time and everyday life (trans. S. Elden and G. Moore). Continuum. C26.P91
Google Scholar Google Preview WorldCat COPAC
Lunberry, C. (2006). Departing landscapes: Morton Feldmanʼs ʻString quartet IIʼ and ʻTriadic memoriesʼ. SubStance 35(2): 17–50. C26.P92
Google Scholar WorldCat

Schmalfeldt, J. (2011). In the process of becoming: Analytic and philosophical perspectives on form in early nineteenth-century C26.P93
music. Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Taylor, B. (2016). The melody of time: Music and temporality in the Romantic era. Oxford University Press. C26.P94

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471563 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

p. 566 Trebay, G. (1983). Voice centerfold: The music of sound. Village Voice, 26 July. http://resoundings.org/PDF/village_voice.pdf C26.P95
Google Scholar WorldCat

Villars, C. (n.d.). Note on the early performance history of Morton Feldmanʼs second string quartet. C26.P96
http://www.cnvill.net/mfsq2perfs.htm
Google Scholar Google Preview WorldCat COPAC

Williams, A., and Srnicek, N. (2014). #Accelerate: Manifesto for accelerationist politics. In R. Mackay and A. Avanessian (eds), C26.P97
#Accelerate: An accelerationist reader, 347–362. Urbanomic.
Google Scholar Google Preview WorldCat COPAC

Wilson, S. (2017). The composition of posthuman bodies. International Journal of Performance Arts and Digital Media (special C26.P98
issue: ʻBodily extensions and performance: Avatars, prosthetics, cyborgs, posthumansʼ) 13(2): 137–152.
WorldCat

Wilson, S. (2018). Notes on Adornoʼs ʻmusical materialʼ during the new materialisms. Music and Letters 99(2): 260–275. C26.P99
Google Scholar WorldCat

Wilson, S. (2021). New Music and the Crises of Materiality: Sounding Bodies and Objects in Late Modernity. Routledge.
Google Scholar Google Preview WorldCat COPAC
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
CHAPTER

27 The Radical Temporality of Drum and Bass  C27

Toby Young

https://doi.org/10.1093/oxfordhb/9780190947279.013.29 Pages 567–C27.P147


Published: 08 December 2021

Abstract
One of the key features of many genres within Electronic Dance Music (EDM) is the creation of
simultaneous temporal layers. Genres such as drum and bass, dubstep, and future bass frequently use
manipulation of rhythmic ostinati and subtle sonic shading to shift the listener’s perception between
these multiple layers; for example, from a fast, intricate motion in the groove, suggestive of the
‘tensed’ experience of A-time, to a slow (or even a-temporal) motion in the vocals, pads, or
instrumental lines, creating a sudden feeling of musical ‘space’, which might in turn connote a
‘tenseless’ B-time. This technique allows producers to create layered temporal narratives within the
music, creating a complex landscape of musical momentum. Drawing on literature and methods from
both sociology and philosophy, this chapter explores the complex relationship between these temporal
systems, and in turn demonstrates how drum and bass o ers a form of temporal resistance to
contemporary life through both the sonic and social experience that the music o ers. It concludes by
arguing that, through the temporal ruptures caused by its uncertain shifting temporality, drum and
bass provides clubgoers with a powerful ontological experience that illuminates the contradictions of
time in a uniquely embodied way.

Keywords: drum and bass, jungle, temporality, rupture, Heidegger


Subject: Philosophy of Music, Music Theory and Analysis, Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online

[Speed] perverts the illusory order of normal perception, the order of arrival of information. What C27.P1
could have seemed simultaneous is diversi ed and decomposes…It is this intervention that
destroys the world as we know it.

(Virilio 1991: 100–101)

Drum and bass is fucking fast. That’s all there really is to it. C27.P2

1
(Goldie)
SINCE the Renaissance, Western modernity has had a profound impact on our experience of time (Koselleck C27.P3
2004) With the ever-increasing rate of social change and technological development brought about by
capitalism, individualism, and scienti c innovation, the project of modernity has created an illusion of a
perpetually increasing tempo of everyday life to the level at which, for many of us, the modern world seems
saturated with the feeling of being in a rush. At an ever-increasing pace of development, digital
technologies are imposing new, faster modes of temporality and rhythm (both social and biological) on our
daily lives, as the media intended to document and evaluate our lives begin to develop the agency to

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
signi cantly alter it. Theorists such as Adrian Mackenzie (2006) have noted that the acceleration of
mechanical (i.e. non-human) temporality through the faster rhythms of data transmission imposed by
digital media splits us from our daily ‘lived’ time, which David Berry (2015) argues imposes a temporal
‘trance’ on users that pushes them out of present time and towards the future.

This rupture between humanity’s daily rhythms and a world pushing forward towards an unknown and C27.P4
unknowable future feels a far cry from the promises of modernity: ‘freedom, progress, [and] in nite human
improvement […] which awaited the possibility of achieving utopian ful lment’ (Scheuerman 2004: 26).
p. 568 Caught in the middle between these colossal forces, we are left experiencing a profound void between
ever-accelerating social temporality and our ontological need to understand the lived presentness of
contemporaneity. Within this narrative of social acceleration, dance music could be seen to o er both a
palliative and an intensi cation of this feeling of temporal inbetweenness. In this chapter, I call attention to
the particular temporal experience within nightclub culture, focusing particularly on the genre of drum and
bass and the unique place it holds in discussions of temporality. Drawing on literature and methods from
both sociology and philosophy, I will demonstrate how drum and bass o ers a form of resistance to
contemporary life through both the sonic and social experience that the music o ers. I will conclude by
arguing that, through the temporal ruptures caused by its uncertain shifting temporality, drum and bass
provides clubgoers with a powerful ontological experience that illuminates the contradictions of time in a
uniquely embodied way.

Club Time C27.S1

As a night-time leisure space, often hidden away in basements or disused post-industrial spaces, the club C27.P5
acts as a form of liminal space set rmly outside both social and biological rhythms, ensuring the
production of a profoundly di erent temporal experience from that of everyday life. Central to this is an
2
immersive EDM (electronic dance music) soundtrack consisting of a succession of repetitive beats;
continuously emitted and seamlessly beat-matched to ensure that a non-stop ow of music—usually at a
similar speed—is achieved across several DJ’s sets throughout the night (sometimes up to 12 hours or
3
longer). Whilst musically pounding and frantic, there’s something about the gathering together of people
outside the normal rhythms of society which invokes a sort of suspension of everyday time. As Mikhail
Bakhtin highlights:

The carnivalesque crowd in the marketplace or in the streets is not merely a crowd. It is the people C27.P6
as a whole, but organized in their own way, the way of the people. It is outside of and contrary to all
existing forms of the coercive socioeconomic and political organization, which is suspended for
the time of the festivity.

(Bakhtin 1984: 255)

The production and consumption of time in a nightclub brings with it complex negotiations of time and C27.P7
space: simultaneously inside yet outside time, public yet intimate, fast yet suspended. Induced by the
sensory overload of intense pulsing beats and ashing light, and perhaps enhanced by drugs or other
stimulants (with the potential to alter biological time even more intensely), the temporal inbetweenness of
the dance oor serves to ‘de- and reconstruct […] any impression of identity, time and location’ (O’Grady
2012: 96); a state of heterochronia, recalling the nowhere and nowhen-ness of the Michel Foucault’s
heterotopia. Such a time of withdrawal from normal modes of social action, notes Victor Turner, can be seen
p. 569 as ‘potentially a period of scrutinization of the central values or axioms of the culture in which it occurs’
(Turner et al. 2017: 167). In other words, being outside a regulated system a ords certain privileges of
experience that allow participants to creatively ‘play out’ and even transcend their daily lived experiences.

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Whether through mechanization of industrialization or the regularity of a typical working o ce day, C27.P8
repetition of mundane cycles forms the basis of temporal demarcation for many (especially in urban
spaces). EDM’s arrival into nightclubs in the 1980s was a rupture to many pre-existing musical conceptions
4
of time, with the relocation of disco’s tools and aesthetics into a repetitive, machinic idiom threatening
‘the ideological and commercial position of white Anglo-American rock that, up to this point, had
dominated the mainstream for several decades’ (Danielsen 2019: 596). Compared with the formalized
linearity of rock, the circular nature of EDM—simultaneously driving yet cyclically hypnotic—initially
posed challenges to popular musicology. There is now a rich body of compelling writing on repetition and
cycles in EDM, which has shifted away from interpretations based on linear/cyclic dualisms and towards
more compelling hybrid models of rhythmic analysis. In particular, work by Christopher Hasty on the
relocation of metre as processual rather than static (1997), Mark Spicer on accumulative form (2004) as a
means of understanding the organic ‘building up’ of a groove as a means for producers and DJs create sonic
narratives, and Mark Butler (2014), whose notion of a ‘perfect loop’—adapted from Robert Fink’s idea of
recombinant teleology—all demonstrate that EDM’s teleological and processual use of looping creates a
space that is far more dynamic and complex than the overly commercialized product of manufacture
capitalism that some earlier scholars took popular music to be (Horkheimer and Adorno 2002).

Despite the ‘relentless four quarter beat’ (Rietveld 1998: 148) that demarcates many of the more C27.P9
mainstream EDM genres like house and techno, Luis-Manuel Garcia (2005) posits that it is not the regular
thuds of the kick drum that drives dance music’s creation and perception but rather the continual stream of
cycles of desire and ful lment. These cycles—perceived primarily by EDM’s characteristic ‘up and down’
energy—are demarcated in the sonic space with indicators of accumulation (such as multitracked build-
ups, drum-roll e ects, and lter sweeps/uplifters) and arrival (the characteristic bass drop). The role of
these aural cues, notes Garcia, is to ‘provide opportunities for listeners to insert themselves into the looping
process and manifest their pro ciency with these massive hypermetrical structures […] by nding ways to
bodily articulate the processes of looping […] mapping out a landscape of shifting creation pleasure while
prolonging the process pleasure of an ever-changing same’ (Garcia 2005: 4.4 and 5.1).

Two Accelerationist Readings C27.S2

Whilst the ways in which dance music’s unique looped temporalities work on musical and perceptual levels C27.P10
p. 570 have been largely agreed upon after over two decades of productive discussion, there is still an ongoing
debate amongst scholars as to where these alternative temporalities t alongside the broader narrative of
social acceleration outlined at the start. On one side is the argument that EDM’s topographical engagement
with industrial urbanity necessarily ties it to the destiny of these spaces (and societies) accelerated futures
(Noys 2014). With a history enmeshed with the spaces torn apart by acts of political or social violence, it was
perhaps inevitable that this would be music that pulled the future towards us in a countercultural
reimagining of the symbolic order of the present. From the sound of its processed drum patterns to the
computerized squelch of Roland’s in uential Tb-303 synthesizer, a strongly machinic aesthetic of intensity
and newness has always permeated EDM; made manifest by an alien DJ estranged from a physicality of
musical labour that was fetishized by most other musical genres.
The central idea from writers such as Benjamin Noys is that we must keep up with the technological and C27.P11
capitalist developments of the commodity economy rather than the impossible or undesirable option of
resisting them. The version of accelerationism known as ‘right accelerationism’ (r/acc)—put forward by
gures like Noys and Nick Land—extends this view by arguing that society must embrace the disruption of
technology to a logical extreme by pushing harder and faster to speed towards a techno-libertarian utopia
of progress and development. The post-humanist agenda behind this certainly speaks to a central tension in
EDM between technology and the body, where the powerful musical prosthesis that technology o ers

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
producers and DJs is in turn used to subvert the ego of dancers into a cyborgian state of simpli ed
consciousness. As MacKenzie Wark comments, ‘the dance does not reveal some aspect of the human, but
rather has the capacity to make the human something else […The] human craves the inauthentic and the
arti cial. This is the basis of its accelerationism: the objective is to encourage machine-made music’s
“despotic drive” of music to subsume both its own past and the presence of the human body’ (Wark 2017).
But this subsumption, the utopian promise of the speed machine, brings with it considerable danger:

[S]peeding-up accelerates us towards the utopian horizon of capitalism, as a social form of ‘pure’ C27.P12
drive and accumulation, ‘freed’ from its dependence on the ‘meat’ of labour—which can then be
read in the ‘communist’ direction of the nal realization of the ‘productive forces’ that shed the
capitalist integument. The thrill also lies in the discarding of the ego, the fusion with the machine
that […] immerses us in capitalist creative destruction. At the same time, we have the ‘threat’ of
obsolescence, social abandonment, and the experience of being condemned to […] dystopian
formulations.

(Noys 2012: 13)

Whilst Noys is rmly in the r/acc faction, his discussion of the existential fear of an ultimately destructive C27.P13
future shows that human obsolescence and burnout is a very real threat for societies undergoing
technological acceleration. Even at the beginning of the last century, commentators had noticed ‘that the
intensity of collective life is too exhausting and must lead to a sort of “individual rest” […with] the rhythmic
p. 571 succession of “high key” and “low key” moments in social life […being] necessary to the very existence
of social life’ (Daynes 2010: 18).

On the other side of this debate, we get an alternative reaction to increasing speed known as left C27.P14
accelerationism (l/acc), which proposes we heed the social exhaustion felt since the last century by ghting
to slow everything down. This form of accelerationism, posited by commentators including Steven Shaviro,
‘seeks to extrapolate the entire globalized neoliberal capitalist order […by] exhaust[ing] it’ in order that
society might be able to reassemble and recuperate (Shaviro 2015: 12). In music, this phenomenon can be
seen in the gradual decrease in the tempi of EDM music in the charts, with the average tempo of songs in the
Club and Dance Top 10 decelerating from 130 bpm in 2010 to 108 bpm in 2017, where it has largely stayed
(Campbell 2017). Whilst speed is not to be confused with acceleration, there are clear entanglements
between them; Rietveld notes that this ‘slow U-turn’ in technoculture (2018: 86) is mirrored at other levels
across the scene, from an increased focus on musical nostalgia in both musical material and themed events
to an increased emphasis on slower bass lines in dance track production, keeping drumbeats further down
in the mix.

Whatever one might make of the validity of these claims of musical deceleration, we see a similar move C27.P15
towards prolongation and di usion in the processual interpretations of Butler and Garcia. The way one
dance track slowly morphs into another o ers a ritualistic experience, which provides a ‘retreat from the
anxieties of time compression and speed, a recurrent moment of redemptive eternity or “mythical” time in
the present […allowing the listener to] turn away from the anxiety of uncertain destiny or an imposing
existential present, to ee into cyclical and repeating patterns of ctional songs, while also stepping outside
accelerating and compressed conceptions of time’ (Negus 2012: 494). In the context of the commodi ed
industry of mainstream popular music that clearly tallies with Adornian fears—churning out ever shorter
songs at a supersonic speed of production and dissemination—the nightclub o ers a di erent sort of non-
mechanical experience, ironically in spite of its mechanicity. Despite clear di erences in felt speed, the
extended musical events of clubbing provide a similar sort of temporal activity to a classical music concert;
existing outside of everyday rhythms and requiring a di erent kind of sustained and engaged attention. As
such, they serve to function as zones of strategic deceleration that a ord participants a way ‘to discipline
themselves and become capable of surviving the onrush of social processes’ (Lilja 2018: 424).

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Entering the Jungle C27.S3

Central to this debate is a fundamental tension between technological acceleration and the human C27.P16
experience of time that we are caught between. Where the machinic desire of r/acc can ‘seem a little
inhuman, as it rips up political cultures, deletes traditions, dissolves subjectivities, and hacks through
security apparatuses’ (Land 1993: 479), the sentimental notion of human primacy in l/acc accounts is
p. 572 troubled by outdated values of humanity. One musical site that might be able to o er something to this
5
discussion is the hyper-intense genre of drum and bass —known in some communities as jungle —whose
sonic identity and musical listening is intrinsically entangled with speed. When the genre exploded onto the
rave scene in the mid-1990s (1992–1997 is often cited as the genre’s prime) against the backdrop of
exponential technological and social change, drum and bass brought with it a radical temporality across
several micro and macro forms that is worth unpacking. With a glitchy and asymmetrical breakbeat at its
heart, this is music whose powerful, central focus on speed pushed the experience of a techno rave to a
liminal extreme. To cater to a growing predilection of ravers for fast tempi, drum and bass sped up the rave,
with beats per minute rising (on average from 130–140 bpm to 170–180 bpm) and the hollow shell of
6
techno’s ‘four-to-the- oor’ giving way to an increasingly playful and active musical surface—a nod to the
growing enthusiasm among ravers for highly syncopated genres such as hip hop and dancehall.

Drum and bass provides a particularly interesting re ection of social temporalities. In their seminal study C27.P17
on class consciousness, John Goldthorpe and David Lockwood posit that working-class communities
towards the end of the twentieth century articulated a ‘present time orientation’, in comparison to a more
middle- (or even upper-) class obsession with the past, manifest in an obsessive focus on maintenance of
value systems, traditional ways of doing things, and cultural objects rooted in history (Goldthorpe and
Lockwood 1963). As with many other genres of dance music—who similarly repurposed the forgotten
factory and dockland spaces of a post-industrialized landscape of post-Thatcherite Britain as their homes
—the intensely ‘in the moment’ experience of rave culture appealed strongly to young, disenfranchised
urbanites, angry at the prejudiced political structures that destabilized communities and restricted social
movement by its citizens. It is not much of a push to see the jittery breakbeat as an ideal vehicle for the
hedonistic speed and forward directionality of the self-perceived underclass. This music also spoke to a
sense of diasporic time for immigrants who were by now used to the ‘relentless e ort of harmonizing the
cultural rhythms of their daily life with the hegemonic temporality of mainstream society’ (Laguerre 2003:
7
6). Drum and bass (and its predecessors, two-step and dub), it has been argued, brought back ‘the
intersection between “fucked up” and “groovy” ’ (Reynolds 2012) which conveyed rave music’s roots in
disco and funk genres.

Erik Davis draws a musical correlation even further back to Black musics, noting an echo in drum and bass C27.P18
of the ‘polyrhythmic sensibility found in traditional West African drumming’ that ‘carves out a unique and
powerful dimension of acoustic space by generating a “nomadic” space of multiplicity that unfolds on the
y’ (Davis 2008: 56). The idea of unfolding is compelling, and is evoked in many of the responses to
interviews conducted with drum and bass participants as part of eldwork in a University of Oxford study.
One participant, Gary, notes, ‘it was like you were in some post-apocalyptic ght, being hyped up by the
pings of sonic warfare and ashing lasers. Man, it was so intense. It just kept driving us to becoming this
8
future army, you know?’ This is backed up by Shanice: ‘as a Black woman, I’d feel like it was my space just
p. 573 like it was other people’s, and we were all there to throw away the baggage of oppression and just party—
9
to become the best, most free versions of ourselves.’ For these communities, the experience of drum and
bass does not align with Lockwood’s present time passivity or hedonism, but instead demonstrated a
politicized and active future-oriented space of exploration and self-creation. Indeed, future-orientation
was stylistically central to drum and bass, as Mark Fisher remarks:

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Jungle sounded like the future rushing in. […] It began as a mutation of rave, con gured from C27.P19
elements of techno, dancehall, dub, rave, house, hip-hop and electro, but the volatile composite
made for a sound that was completely unprecedented: futuristic and experimental, yet designed
for the dance oor […] [B]y fusing elements of science ction and horror, jungle constructed its
own Afro-futurist and cyber-gothic sonic ction, set in the derelict arcades of a near future
populated by millenarian rastas, cyberspace cowboys, voodoo loa and malign shape-shifting
entities.

(Fisher 2011)

Temporality clearly has a big part to play in this aspirational reading: the forward-looking quality of the C27.P20
aural futurism, with its intense, machinic grooves, being matched by a cultural imagination of subaltern
fantasy paradise, played out through the genre’s striking fashion and visual aesthetic. In this alternate
reality, Fisher seems to be hinting at a space where participants could not just be who they wanted to be (as
the PLUR mantra dictated) but even try out the possible futures selves they might wish to inhabit. Whatever
one might make of this reading, the breakbeat’s funkiness clearly reinscribed a Black identity to rave culture
that appealed to a new audience of young Black teenagers (typically second- or third-generation Windrush)
who were drawn in to a world that previously had not spoken to them—allowing them the opportunity to
party and revolt in equal measure. Unlike the suburban or even bucolic quality of the rave scene, drum and
bass was an intrinsically urban culture which re ected its mixed ethnicity at every level of its construction,
from producer to partygoer, whilst maintaining a musical and social uidity that Paul Gilroy suggests
denoted the ‘restless, recombinant qualities’ of an Afro-diasporic culture needing to explore and resist the
spaces of modernity in which it found itself (Gilroy 1993: 31). The ways in which drum and bass allowed the
inscription of minoritized temporalities on the structural and interactional organization of British cities,
and the chance for optimistic self-representation and exploration this inscription o ered young teens with
little socioeconomic hope, really cannot be overstated in any narrative.

Temporal Multiplicities C27.S4

For a genre caught between several time-based multiplicities, it seems inevitable that the breakbeat to C27.P21
p. 574 which drum and bass was anchored was a musical device of both temporal contradiction and potential.
Ever insightful and poetic in his analyses, Fisher describes how:

the jungle breakbeat did not function to keep time. Rather, it multiplied it, twisting and splintering C27.P22
it into an explosion of rhythmically polyperverse foldings and weird angles […] Yet it was also the
intricate involutions of the breakbeats themselves that produced the sense of temporal delirium,
their enfolding recalling the impossible geometries of MC Escher’s work and justifying Kodwo
Eshun’s description of jungle as ‘rhythmic psychedelia’.

(Fisher 2011)
The faster experience of surface movement a orded by drum and bass—an aesthetic described by C27.P23
participants as ‘intricate’—o ered an inhibited freedom to revellers, reinforced by the energizing e ect of
the breakbeat’s internal syncopations. Although still tting into the regular outline of an established and
consistent 4/4 beat pattern with regular pulse, the break typically eschews strong beats in favour of
emphasis on metrically weaker locations (Butler 2006: 79), often changing from one bar to another, making
it harder for a listener to lock in to the groove and creating a feeling that the beat is ‘tripping over itself […as
10
if] missing a step on a ight of stairs’ (Christodoulou 2013: 203). By shifting the listener’s focus from a

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
downbeat to an upbeat, it becomes less clear where to synchronize bodily moment to. However, as Witek
(2017) has noted, adding ‘acoustic energy’ to upbeats in syncopated rhythmic grooves gives impetus to
‘phases in the movement cycle where movements are directed away from gravity’, causing the dancer to ll
in the ‘missing’ beat with more upward and active movements. This feeling gives a more ‘exciting,
propulsive, and light “feel” to the rhythm that is absent in a “straight” rhythm that simply emphasizes the
downbeats’ (Fitch 2016: 6), creating the inspiration and potential for a more expressive range of movement
and even liberated forms of social interaction (Butler 2006). As Witek and colleagues observe:

[T]he syncopations that occur within the loops are equally important in o ering opportunities to C27.P24
create the music. When synchronising our bodies to the beat, we enact parts of the musical
structure by lling in the gaps; as long as the syncopations are repeated, we continue to
participate, and processual pleasure is prolonged. […] There is no drive to resolve tension; rather,
we want to be an active part of it.

(Witek et al. 2014: 151)

Rowan Oliver (2020) and others have observed that an extra level of funkiness and rhythmic interest also C27.P25
comes from the microtiming inherent within the recorded excerpts commonly sampled in dance music—
the most commonly used example in drum and bass being a seven-second lick from a drum solo on the 1969
soul track ‘Amen, Brother’ by The Winstons—which were then manipulated through time-stretching or
digital programming into breaks. As Vijay Iyer elucidates, ‘[because] the original sampled recording bears
the microrhythmic traces of embodiment, the result sounds something like a human drummer improvising
p. 575 with often amusing ourishes and ample metric ambiguity. Momentarily regular, almost human-
sounding pseudo-drumming devolves into inhumanly rapid sequences of rhythmic attacks … [which] serves
to problematize the listener’s image of a human drummer’ (Iyer 2002: 277). This problematization is
particularly evident in the juxtaposition which sits at the heart of dance music between the uidity of a DJ’s
performance in EDM and the seemingly static, typically prerecorded musical objects which are ‘given time’
by processes of repeating, transitioning and transforming (Butler 2014: 71–72).

The choice of samples used in drum and bass animates the continuity of a particularly Black musical C27.P26
lineage, with producers’ utilization of live drum samples such as the Amen break being a way to ‘send
respect to their predecessors’ (Ferrigno 2011: 100) in a virtual, cross-temporal jam with the rhythm sections
11
of bygone eras. This intersection between time, memory, and collective identity invokes not only a respect
of the past but a form of recollection and remembering of periods of Black struggle in digital homage to the
important social function of 1960s and 1970s funk and soul bands. Unlike music produced entirely on
machines, like techno and house, the processing of human drummers through extreme technological
mediation gives drum and bass an ‘uncanny cyborg-like feel’, like a cryogenically frozen body with no
agency of its own. By being accelerated to non-human (or even post-human) speeds, the sampled object has
a tendency towards becoming uncoupled from the rich layers of meaning that the invocation of the past
provides, with samples like the Amen break becoming sorts of hyper-objects that ‘transcends
spatiotemporal speci city through […their] massive and incomprehensible distributions and lifespans’
(Morton 2013: 98).
Shifting from reception to production, the technical ne-tuning required to manipulate these sampled C27.P27
breakbeats advances a further layer of temporal interest. With each individual drum sound painstakingly
manipulated on a sampler or (later) a computer, the long timespan of production that was needed to
facilitate the painstaking micro-programming demanded a new care and attention to time from producers.
Making techno is more about ‘creating a vibe’, notes Joe Baker (Forest Drive West), where the producer just
‘leave[s] it rolling constantly’, and tweaking small details typically through the manipulation of hardware.
In contrast to this fast (and, it is worth noting, cheap) production method, making drum and bass is ‘more

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
stop and start, changing little bits, lots of going back from the beginning of the track, going through and
hearing where an edit needs to be made’ (quoted in Lynch 2011). Drum and bass producers were required to
sit at their computers for days or even weeks of demanding and intense labour. As Simon Reynolds notes, in
a typically caricatured way, ‘you can hear the conditions under which the music came into being: bodies
rigid with tension as they click the mouse; eyes fucked by the red-eye e ects of ganja and staring at a
computer screen all day’ (Reynolds 2017). The uid, in-time ‘virtuosity’ (perhaps sonically, rather than
physically) of techno producers is often taken in the literature to be more impressive and authentic than the
out-of-time craft of temporal manipulation we see in drum and bass. Even Chris Christodoulou, who
usually o ers fairly sympathetic readings of dance music, posits drum and bass as being the product of a
‘fast and cheap’ creative impulse rather than one which demonstrates ‘musical competence in a traditional
p. 576 sense, such as playing in real-time’ (Christodoulou 2011: 52). Within the drum and bass community,
however, temporal endurance is seen as a mark of prestige, as the producer is required to mitigate the
unstoppable ‘velocity of technologico-creative development’ (Monroe 1999: 147) in a ght against the
fallible limits of human concentration.

Dualisms and Limit States C27.S5

Many participants in the scene note a tension between music designed for hedonistic enjoyment ‘in the C27.P28
moment’ yet built on such intense temporal extremes—working at the limits of human physicality—which
meant that fully comprehending it in that moment was almost impossible. The acclaimed drum and bass
producer Goldie comments:

the tunes that really inspire me […have] something about them that’s made for the future. And C27.P29
whether it has impact at the time or impact in the future, the best songs become psalms. They’re
fucking biblical.

(quoted in Jenkins 2019)

Just as in our earlier discussion on accelerationism, the duality of this description—simultaneously C27.P30
invoking a present-day actor looking forward and a retroactive process of canonization taking place in an
imagined future—suggests a musical genre that had sped up so fast as to transcend the structured system of
time entirely; a music so dense in its presentation of sonic (and indeed haptic) material as to render it
simultaneously both mercurial and monolithic. A very practical manifestation of this tension lies in a
widespread bemusement about how to dance to drum and bass. In a subreddit devoted solely to this topic
(‘r/dnbstep’, 2019), members of the community regularly bemoan their di culty in deciding how to
manipulate their bodies to the complex arrangement of rhythmic textures and intense speed of the music:

I was at a Rudimental show not to [sic] long ago and noticed something really strange when they C27.P31
went into some of their harder DnB songs, no one at the show knew how to dance to DnB? When
the beats started hitting hard it looked like the audience was having a massive seizure, not
dancing. No one was moving to the right part of the tune, everyone was trying to dance to the most
frantic part, the snare rolls. News ash you can’t dance to 180 BPM without looking like you are
getting electrocuted. (Fantasimon in ‘r/dnbstep’, 2019)

One of the fun things about DnB is the fact that the bpm is so fast you never know what to do. C27.P32
Everyone says to just do what feels natural but oh my god nothing feels natural when the music is
JUST SO FAST. After a few beers I just give up and become one of those in atable ailing arm
things you see at car dealerships. (hjras in ‘r/dnbstep’, 2019)

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
yessss i love it. skanking, dnb step, x outing or whatever people call it, luuuuuv it:D if im exhausted C27.P33
from 40 min straight dancing like a retard ill just sway back and forth and bang my head from time
p. 577 to time. and if im literally dead, ill just take another one of these pills the nice guy under the
bridge gave me…(lul jk dont take stu from people you dont know).

([user deleted] in ʻr/dnbstepʼ, 2019)

For anyone who has experienced drum and bass dancing rst-hand, these comical images of ailing limbs C27.P34
and self-consciously intoxicated participants concentrating on how to physically parse the extreme speeds
of the music will be all too familiar. With the music’s tempo being so extreme, the spontaneous, non-
technical, and intuitive form dancing by the untrained participant can be described—at its best—as
animating a freely embodied experimentation of the aural-kinesthetics of a social dance space, and—at its
worst—as more of a frantic, un-coordinated desire to simply ‘keep up’ to the music without causing any
physical damage. As a reference to pills suggests (ca eine-based energy drinks are also common as legal
equivalents), many ravers rely on intoxication required in order to increase the dancer’s heart rate to
perform the biological functions needed to match the unnatural speed of the music.

Other posts point to a need to practice movements at home to avoid injury, typically pointing to a few C27.P35
recognized genres of drum-and-bass-speci c practised forms of dancing—including x-outing, skank
stepping—commonly known under the moniker DnB step, but with their roots rmly in hip hop dance-
styles like jumpstyle and the crip walk. These styles revolve around intricate footwork, switching between
heel and toe whilst bouncing back and forth between the two feet, optionally combined with body twists and
other leg-bending manoeuvres. Even for experienced dancers, the sophistication of these movements
requires a genre-speci c expertise and dance experience, in an uncomfortable juxtaposition with the
narrative of ‘doing what feels right’ associated with so much of the PLUR scene. We are perhaps reminded of
Ko Agawu’s discussion of the need for appropriate cultural knowledge in polyrhythmic African music: ‘For
cultural insiders, identifying the gross pulse or the pieds de danse [‘dance feet’] occurs instinctively and
spontaneously, [whereas] those not familiar with the choreographic supplement […] have trouble locating
the main beats and expressing them in movement’ (Agawu 2003: 73). Furthermore, since these specialist
movements and techniques are not necessarily commonly available to participants in the scene apart from
through mimetic learning (although a proliferation of YouTube tutorials on the subject has recently
changed this), the knowledge a ords subcultural capital for the dancer. The unattainability of this
underground knowledge tallies with a recurrent emphasis in the comments above on the helpless or
disabled body; far from the openness and ‘in the moment’ freedom attributed to EDM dancing, participants
often seem genuinely perplexed as to how to condition their body in these di cult temporal conditions.

Simon Reynolds seems to think that the challenges that drum and bass’s extreme speed and rhythmic C27.P36
complexity pose to the dancing body are all but insurmountable: ‘Lagging behind technology, the human
body simply can’t do full justice to the complex of rhythms […] The ideal jungle dancer would be a cross
between a virtuoso drummer (someone able to keep separate rhythms with di erent limbs), a body-
popping breakdance, and a contortionist’ (Reynolds 2017). There is more than a nod to the post-human in
p. 578 the qualities that Reynolds believes are required to negotiate the cybernetic entanglement between
dancer and music, including a seemingly super-cognitive ability to parse the rhythmic complexity of the
music and a cyborgian physicality to maintain the genre’s frenetic dance moves. This narrative of the
super-human requirements of the dancer pervades EDM scholarship but is worth interrogating further.

In music intended to be danced to, it can be expected that modes of listening are mediated and informed by C27.P37
appropriate—or at least acceptable—choreographic practices. When dancing to music in a club, our
attendant modes of listening are guided by choreographic expectations that draw on ‘our internalized
repertoire of already acquired gestures to make sense of a new timing’ (Danielsen 2016: 605; Massumi 1995:

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
67–83). In EDM, the foundation for these movements is an expectation of a regular, isochronous
articulation of time (i.e. a pulse) to which to synchronize the organization of our movements:

[T]he most widespread norm is for dance and music to be coordinated and synchronized around a C27.P38
periodic structure […] While many dances have a clear and formal relationship to the metrical
structure of the accompanying music, demonstrating di erentiated movements for the di erent
beats of the metre, other forms of dancing (and particularly the more informal ones) may show
little or no metrical di erentiation—despite being clearly periodic (i.e. pulse-based). In the
relatively unchoreographed and informal circumstances of club dance music […] bodily
synchronization to periodicity is therefore […a] fundamental consideration.

(Clarke 2016: 109)

The sensorimotor mechanism triggered by group entrainment to a regular pulse is extremely valuable in C27.P39
facilitating feelings of social connection and synchrony between dancers (Krueger 2015). In drum and bass,
however, the hyper-complex rhythmic nature of the music and liminal speed makes the synchronization of
regular body movements to a musical pulse extremely di cult on a practical level (as discussed earlier). An
additional challenge to drum and bass entrainment lies in its plurality of temporal experience, with the
complex polyrhythmic surface and focus on weak beats (rather than downbeats) denying us a clear sonic
pulse on which to focus. As Christodoulou notes, drum and bass’s breakbeats ‘seem to push and pull at the
dancers’ sense of time by suggesting the development of new rhythmic lines. As such their search for a
regular basic pulse generates brief feelings of suspense, tension and even mild panic’ (2013: 203). Whilst
Christodoulou reads this temporal anxiety as in inherent part of the genre’s aesthetic, Reynolds claims that
drum and bass’s ‘body-ba ing’ polyrhythms are only ‘discombobulating unless you xate on and follow
one strand of the groove’ (quoted in Christodoulou 2011: 254), implying that the panicked dancer is simply
not doing it right. Attempting to ‘get around’ the body–mind conundrum of multifaceted groove navigation
by simply ignoring the intrinsic complexity of the music seems like a weak move, not least because the
invitation to negotiate this complex temporal space is what makes dancing to drum and bass so enjoyable.
p. 579
The Silent Disco Paradigm C27.S6

Part of the problem here is that trying to negotiate this tension necessarily invokes an active cognitive C27.P40
element that we simply are not used to engaging with in the deeply instinctual dancing practices typical of
most other EDM genres. Whilst the more ‘obvious’ and ‘natural’ entrained movements of techno and house
are not so readily amenable to conscious re ection and modi cation (Bourdieu 1987), the physical

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
experience of drum and bass clearly needs some thought; yet in most accounts (scholarly and colloquial),
this element is almost always neglected in favour of more corporeal narratives. Perhaps the closest that
scholars have come to unpicking this is Eshun: ‘When polyrhythm phase-shifts into hyper-rhythm, it
becomes unaccountable, compounded, confounding. It scrambles the sensorium, adapts the human into a
“distributed being” strung out across the webbed spider-nets and computational jungles of the digital
diaspora’ (Eshun 1998: 76–77). Where, for Eshun, this quality privileges a post-human concept of the self
who is connected speci cally with technological forces, I suggest that this networked interpretation actually
speaks to a much more human dialectic between mind and body. A moment where this tension becomes
clearly manifest is outlined by Gary:

There’s this moment where you’re dancing and dancing and pushing hard, you know, and C27.P41
suddenly it gets to a point in the night where you have this like epiphany that your mind has
transcended time and space but your body is fucking knackered. <laughs> You suddenly realize
that you can’t give it any more at the full tempo, so you knock back to half tempo and do a little
chilled dance for a while so you can recover before hitting it again and pounding back to full
12
speed.

In this account, the dancer’s sudden awareness of their body’s frailty shifts their attention from the fast C27.P42
speed (seemingly ‘a bit too fast’ for dancing) of the music to a slower-moving element (which might
previously have been thought ‘a bit too slow’ for dancing). Through the pairing of slow, dub bass lines
against fast breakbeats, the juxtapositions of slow and fast tempi make drum and bass feel intrinsically like
13
a multi-speed music; but, as Emma observes, this experience is not just a musical one:

When you need to drop back, it triggers a sort of silent disco in your head—you know those things C27.P43
when you have headphones that play di erent music at the same tempo, but it’s all di erent vibes
so you’re dancing to a down-tempo song but your friend is dancing to an up-tempo song? It’s like
14
you’re icking a switch in your head to make the world slower and more chilled out.

Having our experience become organized di erently by shifting the focus of our attention is commonplace C27.P44
in day-to-day life (see Sartre’s oft-quoted description of his focus shifting from a busy café environment to
his friend Pierre). But key to the aspect change (or gestalt shift) described by Emma is an interruption of the
p. 580 ow state invoked by total immersion of dancing in a club, and the re-interpretation of rhythmic
schemata this causes. Much like Wittgenstein’s famous discussion of Jastrow’s ‘Duck-Rabbit’ (1953: 187–
229), which can be interpreted in two separate ways—either a duck facing left or a rabbit facing right—the
aspect change forces a revelation of something which was not mentally visible, despite always being
optically or aurally present (Noë 2006). Where the clubber was previously immersed in the self-driven
subjectivity of time, the ssure of the aspect change—like the switching of channels in a silent disco—
draws the dancer into a onto-temporal self-awareness, where the relationships of musical events are
objectively drawn into comparison with other external markers of time, such as the limits of human
movement (or other biological functions) or an awareness of clock time induced by thinking about other
activities of the evening (how and when the clubber is going to get home, what they are doing the next day,
etc.).
As the perpetual motion of the body comes grinding to a halt through this cognitive shift, the novelty of C27.P45
inertia intensi es our sense of the present moment in much the same way as becoming jolted awake after
momentarily falling asleep: a sort of biological event that shakes us out of a seemingly out-of-body state
through a rupture in the dominant order of perception. ‘An Event declares that another world is possible,’
remarks Badiou. ‘It ruptures the appearance of normality, and opens a space to rethink reality from the
standpoint of its real basis in inconsistent multiplicity’ (2011: 201). In the visceral club setting, this is
perhaps a more violent and embodied shift than the simple switching of perspective would suggest: more of

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
15
a Zižekian act than a Badiouian event, where the subject is not only shaken out of rhythmic entrainment
but out of the ow state all together. The agency invoked in making this shift (or ‘ icking the switch’)
between temporal schema focuses the dancer’s attention on the fact that, despite being triggered by intense
sonic and visual stimuli, the phenomenological experience of the rhythm they are dancing to is understood
by them as an internal, subjective perception of time. The seeming timelessness o ered by this escape from
the frenetic temporal movement around it mediates both a sense of experienced time and a non-temporal
space to give meaning to the ows of time that the clubber has been experiencing so viscerally up until now.
Zižek comments on this complexity in a typically extravagant passage:

[Inertia] is, of course, not simply outside time; rather, in the ‘stratigraphic’ superimposition, in C27.P46
this moment of stasis, it is time itself which we experience, time as opposed to the evolutionary
ow of things within time. It was Schelling who, following Plato, wrote that time is the image of
eternity - a statement more paradoxical than it may appear. Is time, temporal existence, not the
very opposite of eternity, the domain of decay, of generation and corruption? How then can time be
the image of eternity? […] In short, one should oppose here development within time to the
explosion of time itself: time itself (the in nite virtuality of the transcendental eld of Becoming)
appears within the intra-temporal evolution in the guise of eternity.

(Žižek 2012: 10)

To put some of these ideas back in context, the clubber is not experiencing a stasis outside time, but rather C27.P47
p. 581 an escape of time from within. In responding to music, just as with creating or performing it, clubbers
frequently describe becoming so absorbed that they lose all sense of time (Witek 2019). The rupture of
exhaustion described by a participant on r/dnbstep that draws clubbers out of this ow state, feels,
counterintuitively, as though it is both stopping time (or at least fracturing the constancy of fast movement)
and making the subject aware of (clock) time—a transcendental becoming that inscribes itself into the
perpetual rhythms of the mundane in the guise of a static superimposition, of a crystallization of historical
development. When the body is rested (or a stimulant has kicked in) and the clubber is ready to shift their
attunement out of this stasis and back towards the fast (i.e. double-time) tempo, something interesting
occurs. Raphaël unpacks this moment for us:

People always give up and end up dancing to drum and bass at half speed—you know how trap C27.P48
music sounds really fast? Well, it’s not, but the drums are programmed to sound really fast, giving
it the illusion of speed. Kind of the same concept, but in reverse—but then there’s a point where
you want to go back to the fast speed for a bit, and then what do you do? Most people do a
combination thing—steps and larger body moves in half-time, faster hand moves and quick steps
16
in between—but, I don’t know, it’s just not as satisfying once you’ve broken the ow.

The cognitive framework of Mari Reiss Jones’s (1987) dynamic attending theory shines some light on this C27.P49
phenomenon (see also London 2012). The temporal-perspective model underpinning Jones’s work
describes our process of rhythmic attunement stemming from the separation of auditory stimuli into a
tactus, or ‘referent’ level, from which the perception and anticipation of subdivisions are extrapolated.
According to London, there is an interaction between various levels of metric entrainment surrounding the
tactus, organized by ‘a kind of hierarchic gestalt’ in which they ‘interact with, or even depend on each
other’ (London 2012: 16), led by di erent attentional entrainment mechanisms in the music. At the start of a
drum and bass track, the fast-moving musical surface and genre expectations attune our inner ears to the
faster beat as the referent level, but as the gestalt shift of exhaustion described above occurs, a new ‘lower’
referent level is revealed in the half-time groove. When we re-attend to the original fast-moving music, we
cannot ‘unhear’ this new basic tactus, and so what was previously perceived as a pulse now sounds more
like higher-level subdivision, or even surface detail. Left in a sort of perceptual between-space, we are

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
unsure where to ‘tune into’. This experience of temporal betweenness opens up a phenomenological chiasm
between several levels of cognition, as we nd ourselves at the boundary between activity and passivity,
between thinking and perception, between senses (feeling and hearing), and even within a sense (moving
and being moved). To paraphrase Merleau-Ponty, our understanding of musical movement and time is no
longer just ‘felt from within, but also accessible from without’ in a complex, pluralistic form of cognition
17
that has been opened up to us (Merleau-Ponty 1968: 176).

p. 582
Towards a Temporal Jouissance C27.S7

As Keith Negus observes (2012), music o ers a particularly fertile space to work through this paradox of C27.P50
plurality by attempting to smooth out the tensions and contradictions at play in between our perceptual
levels. In order to trace the phenomenological interactions between these multiple (and con icting) forms
of time and the ways in which society constructs, frames and shares its lived experiences, Jacques Lacan
(1966) examines the relationship between a subjective experience of duration and the objecti cation of this
18
experience in the well-known thought-experiment of the prisoner’s dilemma. Applying the apparatus of
game theory, Lacan attempts to subvert the phenomenological distinctions between objective ‘time’ and
subjective ‘temporality’ by trying to explain a way in which the past and future interact at the ‘edges’ of the
present in order to ‘disrupt the assumed synchronicity of logic’ (Lapping 2017: 9). He concludes that rather
than superimposing the past and future on the present, an intrasubjective view of time occurs in three
moments: the instant for seeing (retroaction), the time for understanding (present) and the moment of
concluding (anticipation). These moments, or modes of duration that modulate the unity of ‘a certain time’,
exist within fundamental tensions between waiting for the past, hesitation at the moment of the present,
and a moment of concluding that ‘is arrived at in haste, in anticipation of future certainty’ (Lacan 1966:
209). Whilst Lacanian theory is surprisingly uncommon in drum and bass scholarship, the interplay
between Lacan’s multiplicity of temporal registers as they are brought together in a single moment
resonates with a short but powerful passage from Davis:

[T]he jarring hyperspace of jungle arises at least as much from the music’s almost eschatological C27.P51
polyrhythm, its deployment of ‘heterogeneous blocks of space-time’ that cut across the
conventional dimensions of acoustic space. These blocks pull compress that space through
intensity and speed, creating little black holes of multiple beats.

(Davis 2008: 69)

The increased density produced by the compacting of time-space into black holes is a powerful metaphor C27.P52
for the weight of temporal con ict that is brought upon the drum and bass listener’s symbolic shoulders as
they experience the intense jouissance of the collusion of time. As Claudia Lapping observes, for Lacan, ‘the
bizarre turns and shifts in our taken for granted experience of “time”, of duration and of urgency […] reveal
not only the fantasmatic status of temporality, but also the shift itself as a point of impact of something else
[…The] juxtaposition of bizarrely incongruent temporalities […] will produce ever more traumatic
registrations of horri c, overwhelming and unknowable Real Time’ (Lapping 2017: 12). In other words, the
simultaneity caused by compression or superimposition of these three temporalities open up interstices or
ruptures in the fabric of our experience, which has the potential to reveal something new about our
p. 583 experience of temporal inhabiting. Any bends, junctures, or shifts in the topological surface of time cause
it stress, bruising time’s surface in a way that reveals something new about our experience of temporal
inhabiting.

When dancing to drum and bass, we perceive the disparate temporal elements together as part of one whole, C27.P53
our fragmented experiences of worldly temporality momentarily held together in our consciousness
19

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
through the unity of the musical work. As we brie y stand outside time, disconnected from the dense and
narcissistic layers of our identity on which we normally focus, we have the potential to reconnect with
something more primordial and essential lying beyond our everyday experience. At this juncture, we are
a orded through this entanglement of perception and cognition an embodied yet unpronounceable
understanding of what being in the world—being truly present in the here and now—might actually feel
like. Such an approach resonates strongly with the early writings of Martin Heidegger. In Heidegger’s own
words, ‘the ecstatical unity of temporality—that is, the unity of the […] raptures of the future, of what has
been, and of the present—is the condition for the possibility that there can be an entity which exists as its
“there” ’ (Heidegger 1967: 350).

For Heidegger, a central role of Dasein is in the revealing (and indeed concealing) of being through the C27.P54
considered thinking and engagement with the external functions of selfhood and society. Heidegger
frequently returns to the image of the Dasein as a thinker, who—through an engagement with the framing
function of the word—‘might be admitted more and sooner and ever more primally to the essence of that
which is unconcealed and to its unconcealment, in order that he might experience as his essence his needed
belonging to revealing’ (Heidegger 1967: 26). This understanding, notes Heidegger, is focused primarily on
the future, in a constant state of anticipation of the inevitable conclusion of life. As a being-towards-death,
the subject is constantly projecting forward towards its end (a concept known as zukommen, ‘coming
towards’) rather than being able to inhabit the present. As he explains:

[O]ne’s state-of-mind temporalizes itself as a future which is ‘making present’. And all the same, C27.P55
the Present ‘leaps away’ from a future that is in the process of having been, or else it is held on to
by such a future. Thus we can see that in every ecstasis, temporality temporalizes itself as a whole;
and this means that in the ecstatical unity with which temporality has fully temporalized itself
currently, is grounded the totality of the structural whole of existence […] Temporalizing does not
signify that ecstases come in a ‘succession’ […but] as a future which makes present in the process
of having been.

(Heidegger 1967: 401)

The experience of temporal collapse felt by a clubber embedded in a drum and bass night compresses all C27.P56
three modes of time together in a phenomenological breach. This powerful moment of temporal rupture
brings Dasein into a rare encounter with the present, and ultimately, momentarily distracts the being-
towards-death from its future-facing orientation. The Dasein is free not only from its future but also from
the past—the personal and cultural baggage of ‘having-been-ness’ (Gewesenheit) that Heidegger believes
20
p. 584 constantly mediates our expectations and assumptions of the future —in an intense form of temporal
release from the constant burden of movement and death. The transcendental joy induced by this sense of
release is extraordinary. The rupture of this temporal release, with its suspension of the normality of our
everyday familiarity of existence, discloses our unique being-there. Much like Lacan’s jouissance, the
intense feeling for the subject as it approaches a state of pleasure serves to reveal something about the
world that was not previously understood. But, in a key point of di erence between the two thinkers, where
Lacan’s revelation is through unlocking the unconscious mind, for Heidegger (Dahlstrom 2000), an
exegesis of truth can only occur when the ‘original truth’ within the world marries up with states of a airs
in which it is inscribed, and reveals itself. As James Luchte notes, for Heidegger ‘we can only know our own
self when it has been resisted, broken or has encountered a limit-situation, via which each nds herself in
her “truth” ’ (Luchte 2011: 49). Through its encounter with an immanent temporal destruction, the excess
of ecstatic feeling evoked is surely a breakage strong enough to reveal this truth.

Revolutionary Time(s)? C27.S8

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
There is one last challenge to the Heideggerian reading that we must navigate. Transcending the discourse C27.P57
of the subject, the encounter between Dasein and the unconcealed truth necessarily occurs in a manner
which is both beyond subjectivity and deeply embodied within the world (Mulhall 2005). Positioning the
worldliness of this unconcealment with the seemingly dislocated subject of Dasein therefore poses a
problem to Heidegger scholars (Carman 2003), who have to decide whether to defend truth-as-
unconcealment as a fundamental paradox, or to consider it to be a more general condition for the possibility
of another form of worldly truth, grounded in interaction rather than individuation, that might be revealed
at a later stage. To conclude this chapter, it seems appropriate to re ect on the highly social interactions at
the heart of EDM culture and consider the validity of a reading of ecstatic temporality in drum and bass
which (re)frames a Heideggerian view of temporality through the lens of dance music’s sociality.

For the Heidegger of Being and Time, the nitude of a being-towards-death is located only within the realm C27.P58
of the individual who learns that they inevitably face death alone: ‘Being-with existentially determines
Dasein even when an other is not factically present and perceived’ (Heidegger 1967: 113). Yet Dasein, in an
anxious attempt to understand its own mortality, reaches to society for the tools by which to measure and
understand its forward-facingness. We might typically think of these tools as being the processes of
temporal documentation and measurement that have become entangled in our contemporary world—from
the social networks that ensnare most of our waking hours to the wearable technologies that monitor our
health constantly. Indeed, provoked by this inherent need to document, we habitually submit extensive
amounts of personal data to companies like Amazon, Apple, Facebook, and Google in order to provide them
p. 585 —after some algorithmic extrapolation—with a time-coded account of where we are, who we are with,
and how we feel, as a way of ensuring the measurements literally are ‘to hand’ in our phones. But for in
spite documentation of the literal present, these technologies do nothing to establish the ‘presentness’ or
contemporaneity of the continually teleological situation in which the subject nds itself. Perhaps we might
21
then think of a tool in a more Heideggerian way (i.e. das Zeug):

Equipment can genuinely show itself only in dealings cut to its own measure (hammering with a C27.P59
hammer, for example); but in such dealings an entity of this kind is not grasped thematically as an
occurring Thing, nor is the equipment-structure known as such even in the using. The hammering
does not simply have knowledge about the hammer’s character as equipment, but it has
appropriated this equipment in a way which could not possibly be more suitable.

(Heidegger 1967: 69)

For Heidegger, a tool is ‘the primordial modality of existence in which we are integrally involved with our C27.P60
world’ (Thomson 2011: 83). Logically, to achieve an ecstatic temporal experience requires a tool that forces
us to acknowledge our relationship with the present (through the coming together of temporal modes (as
discussed earlier), not by documenting our presentness, but by drawing our attention back to the present via
the ‘thickness’ of temporal transcendence. The only logical candidate, in a state of sensual engagement with
the music surrounding it, is the dancing body. It does not seem enough, though, to suppose that a body,
even one in motion, might o er the appropriate ‘equipmentality’ by itself; rather, it will surely only o er up
its ‘equipmentality’ in a situation that demands the tool be used in such a way as it becomes a tool
(Thomson 2011: 82).
For the body to be able to facilitate the experience of temporal ecstasy, it needs to be open to experiences of C27.P61
time, both internal and external. It needs to become ‘a membrane open to the outside […] in between
dimensions’ (Fernández 2007: 83) with an ‘expandable and pliable’ nature (Massumi 2002: 231) to facilitate
the necessary meeting of outside and inside. Embodiment of the machinic patterns of dance music is not
enough, therefore, without the groundwork of social interaction and experience attached to the clubbing
experience: its role as a night-time leisure activity (outside everyday rhythms), the multisensory stimuli of
strobe lighting, temperature, visual e ects, and the imbibing of hallucinatory drugs or other stimulants. By

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
going to these spaces, participants are required not only to lower their social barriers but to make
themselves open to achieving alterity. Whilst also facilitating the relational blurring between dancers’
individual subjectivities, lowering these barriers forces clubbers to be open to vulnerability (in a way that
only uninhibited movement in front of strangers can be) in order to make themselves available to a greater
condition of embodied harmony or ‘perfect contingency’ with their physical surroundings (Schüll 2012).

Invoking bodily vulnerability is radical in a society where biopolitical freedom is at a premium (Sharma C27.P62
2014). The lack of control (or even autonomy) over their time that the condition of social acceleration
a ords disenfranchised subjects—unless of course they try to ‘catch up’ by increasing productivity—
p. 586 demonstrates a society locked within a damaging neoliberal subjectivity hurtling towards a future that it
cannot provide. Rather than being governed by individual interpretation of the ebb and ow of personal
time, this form of time is driven by the brutal alignment of bodies within the ‘turbo capitalism’ that fuels it
(Han 2017). We might come close to feeling the di erent temporal registers of transcendent unpredictability
around us, but temporality that resists compliance with this social ideology, that does not attempt to
‘achieve’ or ‘strive’, as predominant models of temporality say it should—typically racialized, gendered, or
queered temporalities, for instance—is considered deviant.

The primary function of drum and bass’s revelation of temporal ecstasy then becomes incredibly apparent: C27.P63
helping those without access to the mechanisms of temporal critique to make visible the entangled and
uneven politics of temporality. Not only do the perceptual ruptures alter time for the clubber, but the
revelation of temporal agency that the experience a ords has the potential to inspire and enable
participants to reclaim control of their own lived time and push against the incarceration of social and
political acceleration in which they nd themselves. In going beyond the everyday limits of human
experience and engaging with time on deeper levels of a ect and subjectivity, the clubber’s experience of
what it means to be human is transformed: inspired to break out of the linear streams of this neoliberal
temporality, their renewed understanding and expanded perception enable them to challenge and
deterritorialize the dominant narratives of time around them.
Conclusions C27.S9

The narrative that drum and bass’s intense speed helps participants to manage the ‘bewildering experience’ C27.P64
of social acceleration (Christodoulou 2013; 2011) by reinforcing, refracting, and challenging dominant
modes of temporality is not a new one in EDM scholarship. Many leading gures in the dance community
have made claims about the relationship between the out-of-time suspension of accelerating social

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
rhythms felt by dancers in a club environment and localized feelings of ‘suspension of embodied
acceleration’ (Gálvez 2019); but previous work has largely avoided a more thorough interrogation of the
implications of these claims and the mechanisms through which these processes unfold. Through a
discussion of the confrontation with quotidian time invoked by a heightened dissociative- ow state and
strongly embodied perception of speed, this chapter has shown how drum and bass forces us to confront the
speeding up of society in a totally new way–not by trying to delineate but by trying to enmesh. By o ering a
phenomenological response to the duality of accelerationism (discussed above) through an ecstatic
experience of the simultaneous ‘thickness’ of a squashed present, drum and bass o ers a powerful aesthetic
and epistemological apparatus for critically delineating, authenticating, and remaking lived experiences of
time in the modern world.

p. 587 In addition, this chapter has considered how, in the highly social space of a club setting, the temporal C27.P65
complexities, juxtapositions, and paradoxes of drum and bass questions these received models by opening
up internal and interpersonal conduits that allow us—even force us—to engage in a deeply embodied
ontology of time. Taking a Heideggerian approach has demonstrated that the ontologically forward-
facingness of our experience as a being-towards-death can be mitigated (if never reversed) by the temporal
ecstasy triggered by an extreme limit state of bodily exhaustion experienced when dancing to drum and
bass. In this Heideggerian reading, forcing the subject to (re)construct time in a new and pluralistic way lies
at the heart of what is so extraordinary about the temporal experience of this genre: in other words, to
answer my own question from earlier, it shows us that the ideal mode of interaction with dance music is one
which, rather than re ect modern time, collapses it. With a perceptual multiplicity opened to us, the
shifting boundaries and manifold rhythmic pathways force even the greatest dancer to confront a complex
space of shared di erence and possibility, in a clear metaphor for our everyday social interaction. As Davis
notes, in a typically optimistic manner, drum and bass provides:

[an] unusually intuitive avenue, not just to conceptualize, but to train ourselves to cross and C27.P66
combine heterogeneous spaces, chaotic ruptures, and zones of communication directly into our
body–minds. As we weave ourselves into polyrhythm’s brillating tapestry of molecular beats and
percussive patterns, we taste a kind of wisdom, not the re ection in repose we conventionally
imagine, but the wisdom of networked di erences, whether social, ecological, or spiritual. In
humanistic terms, we are shown how the action of standing apart from others can actually support
the entire group, and how novelty is not so much an imposition of creative individual will as a kind
of active remix of other beings and energies.

(Davis 2008: 59)

Maintaining this social space (both literally and theoretically) in discussions of contemporary musical time C27.P67
is vital in negotiating our relationship with the rich tapestry of temporalities and states of time that we are
so keen to demarcate and document in other areas of our lives. As compelling and ‘neat’ as more techno-
deterministic readings of experiences EDM time of scholars like Danielsen might be, Sarah Thornton
reminds us ‘not to deprive clubbers and ravers of their agency […as] active and creative participants in the
formation of club cultures’ (Thornton 1995: 244). In other words, the physico-ontological reading of drum
and bass put forward in this chapter is necessarily a human creation, and can only be achieved through a
collaborative openness between clubbers through which a temporal ecstasy might be allowed to reveal itself.
Such an experience holds an immense potential to reveal not only a di erent way of being outside the
dominant and controlling model of state apparatus but how to become mobilized for political activity and
change. It is an incitement for us to be open to the radical humanizing of the powerful temporal experiences
we encounter in the world around us—as well as the vulnerability these experiences necessarily arouse—
that make us confront our complex relationships with temporal paradoxes. It is a plea to embrace the
p. 588 dualities between machine/human and body/mind that invoked through drum and bass, help us open our
eyes to the temporal constraints and imbalances of our daily lives. And most of all, it is a cry to

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
deterritorialize the oppressive ows of society and invite us to rebelliously and joyfully explore the
experiences and implications of temporal di erence and otherness within a shared subjectivity.

Notes
1. Pre-concert talk (Southbank Centre, London), 22 Oct. 2017. C27.N1

2. Whilst EDM is broadly conceived to include a variety of subgenres and scenes, the discussion is largely focused on genres C27.N2
that are intended to be danced to (i.e. not glitch or IDM), and sit within the median range of music typically played in club
settings (i.e. not gabber, speedcore, extreme psytrance at the fast end, or ambient, industrial, and some dubstep at the
other).

3. This chapter specifically considers the temporality of night-long club and rave events, rather than shorter sets in EDM C27.N3
festivals, which are normally no more than one hour long. The formation of alternate temporal spaces does occur in these
festivals—albeit in slightly di erent ways—as discussed in work of Graham St John (2019) and others on the ʻweekend
societiesʼ that gather together for EDM music festivals like Burning Man in Nevada, USA.

4. A proclivity of drum machines and DAWs (digital audio workstations) as the primary site through which EDM was C27.N4
produced also brought an ideological challenge to the primacy of played rather than machine-generated performance, as
discussed at length by Anne Danielsen (2019).

5. Whilst there is an ongoing discussion in the drum and bass community about whether or not the two terms di er, they are C27.N5
used synonymously here. For anyone unfamiliar with the drum and bass idiom, some examples that capture the rhythmic
play and temporal subversion of this discussion particularly well are: ʻJungle Crackʼ (2016) by London-based producer
Forest Drive West, ʻHeadroomʼ (Audio VIP) (2013) by collaborative production group Cause4Concern, and ʻSeven Samuraiʼ
(1998) by influential LA-based artist Photek.

6. The regular 4/4, on-the-beat kick and snare patterns of house and techno. C27.N6

7. It is worth noting that this was not always the case: despite the pervasive narrative of inclusivity or PLUR (ʻpeace, love, C27.N7
unity, and respectʼ) that permeates scholarship on EDM culture, many of the dance genres that preceded drum and bass
were somewhat uniform in their temporal experience. Techno, for example, in pursuit of a proto-communist
egalitarianism, was o en guilty of whitewashing the temporal expereince of participants into a beat as industrial and
regulated as the warehouses it was in.

8. Interview with Gary, 4 June 2019; conducted as part of fieldwork undertaken at the University of Oxford alongside the C27.N8
public engagement project ʻRemixing the Genreʼ.

9. Interview with Shanice, 12 Mar. 2019. C27.N9

10. Danielsen also notes the importance of sonic dislocation of the genre in creating this e ect: ʻit is not first and foremost the C27.N10
transformation of temporal features or durations that produce the peculiar microrhythmic e ects [in DnB] but the cutting
up of sounds and the abrupt transitions between sounds that such cuts produceʼ (2019: 599, emphasis original).

11. See Stanyek and Piekutʼs (2010) work for a further discussion of this sort of ʻintermundane collaborationʼ. C27.N11

p. 589 12. Interview with Gary, 4 June 2019. C27.N12

13. Iyer (1998) notes that ʻthe discernment of qualities […] from a piece of music is not perceptually inevitable; rather, the C27.N13
music may o er perceptual ambiguities whose resolution depends on an observerʼs culturally contingent listening
strategiesʼ (pp. 83–104). In particular, a recent study by Mari Reiss Jones (2019) demonstrated that, as people age, they
tune in to increasingly slower sets of privileged event-based driving rhythms. All of the interviewees quoted in this chapter
were under the age of 40, but further work with a wider demographic of interviewees would be valuable to avoid
privileging the musical perception of young participants over that of older clubbers.

14. Interview with Emma, 23 Mar. 2019. C27.N14

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
15. For Zižek, the fracture of the act causes a break from the ʻsymbolic coordinatesʼ in which it occurs, without hope of C27.N15
restoration. It is not so much a revelation, as both a damaging of the subject (compare here with the body ʻbrokenʼ from
its dancing by excess and exhaustion) and a ʻself-wounding of subjectivity itself […which] touches the real— the gap or
point of failure within any symbolic structure—and leaves the subject without the network of support that sustains a
sense of meaningʼ (McGowan 2010: 10).

16. Interview with Raphaël, 1 Apr. 2019. C27.N16

17. From an account of the experience of one hand touching the other, and the phenomenological limit-space between C27.N17
activity and passivity that this act creates.

18. For a more expanded account, see Lapping (2017). C27.N18

19. The word ʻecstasyʼ is etymologically derived from ex (apart) and stasis (standing). C27.N19

20. Note that Heidegger does not believe this to be a negative thing, but rather allows the subject to take better control over C27.N20
the future with the increased knowledge of the worldʼs framework a orded by the past—a phenomenon he calls
ʻresolutenessʼ (Entschlossenheit).

21. Also sometimes translated as ʻequipmentʼ. C27.N21


References C27.S10

Agawu, K. (2003). Representing African music: Postcolonial notes, queries, positions. Routledge. C27.P68
Google Scholar Google Preview WorldCat COPAC

Badiou, A. (2011). Being and event. Continuum. C27.P69


Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Bakhtin, M. M. (1984). Rabelais and his world. Indiana University Press. C27.P70
Google Scholar Google Preview WorldCat COPAC

Berry, D. M. (2015). Critical theory and the digital. Bloomsbury Academic. C27.P71
Google Scholar Google Preview WorldCat COPAC

Bourdieu: (1987). What makes a social class? On the theoretical and practical existence of groups. Berkeley Journal of Sociology C27.P72
32: 1–17.
Google Scholar WorldCat

Butler, M. (2006). Unlocking the groove: Rhythm, meter, and musical design in electronic dance music. Indiana University Press. C27.P73
Google Scholar Google Preview WorldCat COPAC

Butler, M. J. (2014). Playing with something that runs: Technology, improvisation, and composition in DJ and laptop performance. C27.P74
Oxford University Press.
Google Scholar Google Preview WorldCat COPAC

Campbell, M. (2017). The changing BPMs of club and dance music. https://www.playnetwork.com/blog/2017/03/13/the- C27.P75
changing-bpms-of-club-dance-music/
WorldCat

Carman, T. (2003). Heideggerʼs analytic: Interpretation, discourse and authenticity in ʻBeing and timeʼ. Cambridge University Press. C27.P76
Google Scholar Google Preview WorldCat COPAC

Christodoulou, C. (2011). Rumble in the jungle: City, place and uncanny bass. Dancecult 3: 44–63. https://doi.org/10.12801/1947- C27.P77
5403.2011.03.01.03
Google Scholar WorldCat

p. 590 Christodoulou, C. (2013). DJs and the aesthetic of acceleration in drum ʻnʼ bass. In B. A. Attias, A. Gavanas, and H. C. Rietveld C27.P78
(eds), DJ culture in the mix: Power, technology, and social change in electronic dance music, 195–218. Bloomsbury Academic.
Google Scholar Google Preview WorldCat COPAC

Clarke, E. (2016). Rhythm/body/motion: Trickyʼs contradictory dance music. In A. Danielsen (ed.), Musical rhythm in the age of C27.P79
digital reproduction, 105–120. Routledge.
Google Scholar Google Preview WorldCat COPAC

Dahlstrom, D. O. (2000). Heideggerʼs concept of truth. Cambridge University Press. C27.P80


Google Scholar Google Preview WorldCat COPAC

Danielsen, A. (ed.) (2016). Musical rhythm in the age of digital reproduction. Routledge. C27.P81
Google Scholar Google Preview WorldCat COPAC

Danielsen, A., 2019. Glitched and warped: Transformations of rhythm in the age of the digital audio workstation. In C27.P82
M. Grimshaw-Aagaard, M. Walther-Hansen, and M. Knakkergaard (eds), The Oxford handbook of sound and imagination, vol. 2,
593–609. Oxford University Press. https://doi.org/10.1093/oxfordhb/9780190460242.013.27
Google Scholar Google Preview WorldCat COPAC
Davis, E. (2008). ʻRoots and wiresʼ remix: Polyrhythmic tricks and the Black electronic. In P. D. Miller (ed.), Sound unbound: C27.P83
Sampling digital music and culture, 53–72. MIT Press.
Google Scholar Google Preview WorldCat COPAC

Daynes, S. (2010). Time and memory in reggae music: The politics of hope, music and society. Manchester University Press. C27.P84
Google Scholar Google Preview WorldCat COPAC

Eshun, K. (1998). More brilliant than the sun: Adventures in sonic fiction. Quartet Books. C27.P85

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Fernández, M. (2007). Illuminating embodiment: Rafael Lozano-Hemmerʼs relational architectures. Architectural Design 77: 78– C27.P86
87. https://doi.org/10.1002/ad.490
Google Scholar WorldCat

Ferrigno, E. D. (2011). The dark side: representing science fiction in drum ʻnʼ bass. New Review of Film and Television Studies 9(1): C27.P87
95–104. https://doi.org/10.1080/17400309.2011.521722
Google Scholar WorldCat

Fisher, M. (2011). The 20 best jungle records ever made. https://www.factmag.com/2011/02/12/20-best-jungle/ (accessed C27.P88
2.22.20).
WorldCat

Fitch, W. T. (2016). Dance, music, meter and groove: A forgotten partnership. Frontiers in Human Neuroscience 10(64): 1–7. C27.P89
https://doi.org/10.3389/fnhum.2016.00064
Google Scholar WorldCat

Gálvez, J. (2019). On analyzing EDM DJ sets: Problems and perspectives for a sociology of sound. In M. Dumnić Vilotijević and C27.P90
I. Medić (eds), Contemporary popular music studies, 149–159. Springer Fachmedien. https://doi.org/10.1007/978-3-658-25253-
3_14
Google Scholar Google Preview WorldCat COPAC

Garcia, L.-M. (2005). On and on: Repetition as process and pleasure in electronic dance music. Music Theory Online 11(4). C27.P91
https://mtosmt.org/issues/mto.05.11.4/mto.05.11.4.garcia.html
Google Scholar WorldCat

Gilroy, P. (1993). The Black Atlantic: Modernity and double consciousness. Verso. C27.P92
Google Scholar Google Preview WorldCat COPAC

Goldthorpe, J. H., and Lockwood, D. (1963). A luence and the British class structure. Sociological Review 11(2): 133–163. C27.P93
Google Scholar WorldCat

Han, B.-C. (2017). The agony of Eros. MIT Press. C27.P94


Google Scholar Google Preview WorldCat COPAC

Hasty, C. (1997). Meter as rhythm. Oxford University Press. C27.P95


Google Scholar Google Preview WorldCat COPAC

Heidegger, M. (1967). Being and time. Blackwell. C27.P96


Google Scholar Google Preview WorldCat COPAC

Horkheimer, M., and Adorno, T. W. (2002). Dialectic of enlightenment. Stanford University Press. C27.P97
Google Scholar Google Preview WorldCat COPAC

Iyer, V. (1998). Microstructures of feel, macrostructures of sound: Embodied cognition in West African and African-American musics. C27.P98
Doctoral dissertation, University of California.
Google Scholar Google Preview WorldCat COPAC

Iyer, V. (2002). Embodied mind, situated cognition, and expressive microtiming in African-American music. Music Perception C27.P99
19(3): 387–414. https://doi.org/10.1525/mp.2002.19.3.387
Google Scholar WorldCat

Jenkins, D. (2019). Goldie interview: Past, present and future of drumʼnʼbass.https://www.redbull.com/gb-en/goldie-drum-and- C27.P100
bass-genre-reflections

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Jones, M. R. (1987). Dynamic pattern structure in music: Recent theory and research. Perception and Psychophysics 41(6): 621– C27.P101
634. https://doi.org/10.3758/BF03210494
Google Scholar WorldCat

Jones, M. R. (2019). Time will tell: A theory of dynamic attending. Oxford University Press. C27.P102
Google Scholar Google Preview WorldCat COPAC

p. 591 Koselleck, R. (2004). Futures past: On the semantics of historical time. Columbia University Press. C27.P103
Google Scholar Google Preview WorldCat COPAC

Krueger, J. (2015). The a ective ʻweʼ: Self-regulation and shared emotions. In T. Szanto and D. Moran (eds), The phenomenology C27.P104
of sociality: Discovering the ʻweʼ, 263–277. Routledge.
Google Scholar Google Preview WorldCat COPAC

Lacan, J. (1966). Écrits. Routledge. C27.P105


Google Scholar Google Preview WorldCat COPAC

Laguerre, M. S. (2003). Transglobality and diasporic temporality. In M. S. Laguerre (ed.), Urban multiculturalism and globalization C27.P106
in New York city: An analysis of diasporic temporalities, 6–27. Palgrave Macmillan. https://doi.org/10.1057/9780230503748_2
Google Scholar Google Preview WorldCat COPAC

Land, N. (1993). Making it with death: Remarks on thanatos and desiring-production. Journal of the British Society for C27.P107
Phenomenology 24(1): 66–76. https://doi.org/10.1080/00071773.1993.11644272

Lapping, C. (2017). The explosion of real time and the structural conditions of temporality in a society of control: durations and C27.P108
urgencies of academic research. Discourse 38(6): 906–922. https://doi.org/10.1080/01596306.2016.1185092
Google Scholar WorldCat

Lilja, M. (2018). The politics of time and temporality in Foucaultʼs theorisation of resistance: Ruptures, time-lags and C27.P109
decelerations. Journal of Political Power 11(3): 419–432. https://doi.org/10.1080/2158379X.2018.1523319
Google Scholar WorldCat

London, J. (2012). Three things linguists need to know about rhythm and time in music. Empirical Musicology Review 7(1–2): 5– C27.P110
11. https://doi.org/10.18061/1811/52973
Google Scholar WorldCat

Luchte, J. (2011). Heideggerʼs early philosophy: The phenomenology of ecstatic temporality. Continuum. C27.P111
Google Scholar Google Preview WorldCat COPAC

Lynch, W. (2011). Breaking through: Forest Drive West. https://www.residentadvisor.net/features/3349 C27.P112


WorldCat

MacKenzie, A. (2006). Transductions: Bodies and machines at speed. A. & C. Black. C27.P113
Google Scholar Google Preview WorldCat COPAC
Massumi, B. (1995). The autonomy of a ect. Cultural Critique 31 (Autumn): 83–109. https://doi.org/10.2307/1354446 C27.P114
Google Scholar WorldCat

Massumi, B. (2002). Parables for the virtual: Movement, a ect, sensation. Duke University Press. C27.P115
Google Scholar Google Preview WorldCat COPAC

McGowan, T. (2010). Subject of the event, subject of the act: The di erence between Badiouʼs and Žižekʼs systems of philosophy. C27.P116
Subjectivity 3(1): 7–30. https://doi.org/10.1057/sub.2009.31

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Google Scholar WorldCat

Merleau-Ponty, M. (1968). The visible and the invisible (trans. A. Lingis). Northwestern University Press. C27.P117
Google Scholar Google Preview WorldCat COPAC

Monroe, A. (1999). Thinking about mutation: Genres in 1990s electronica. In A. Blake (ed.), Living through pop, 146–158. C27.P118
Routledge. https://doi.org/10.4324/9780203012734-17
Google Scholar Google Preview WorldCat COPAC

Morton, T. (2013). Realist magic: Objects, ontology, causality. Open Humanities Press. C27.P119
https://doi.org/10.3998/ohp.13106496.0001.001
Google Scholar Google Preview WorldCat COPAC

Mulhall, S. (2005). Heidegger and ʻBeing and timeʼ (2nd edn). Routledge. C27.P120
Google Scholar Google Preview WorldCat COPAC

Negus, K. (2012). Narrative time and the popular song. Popular Music and Society 35(4): 483–500. C27.P121
https://doi.org/10.1080/03007766.2011.567918
Google Scholar WorldCat

Noë, A. (2006). Action in perception. MIT Press. C27.P122


Google Scholar Google Preview WorldCat COPAC

Noys, B. (2012). Speed machines. Nyx 7: 10–18. C27.P123


Google Scholar WorldCat

Noys, B. (2014). Malign velocities: Accelerationism and capitalism. Zero Books. C27.P124
Google Scholar Google Preview WorldCat COPAC

OʼGrady, A. (2012). Spaces of play: The spatial dimensions of underground club culture and locating the subjunctive. Dancecult C27.P125
4: 86–106. https://doi.org/10.12801/1947-5403.2012.04.01.04
Google Scholar WorldCat

Oliver, R. (2020). Breakbeat syncretism: The drum sample in African American popular music. https://www.reddit.com/r/dnbstep/ C27.P126
Google Scholar Google Preview WorldCat COPAC

p. 592 Reynolds, S. (2012). Energy flash: A journey through rave music and dance culture. So Skull Press. (Originally published as C27.P127
Generation Ecstasy, Little, Brown, 1999.)
Google Scholar Google Preview WorldCat COPAC

Reynolds, S. (2017). The Wire 300: Simon Reynolds on the Hardcore Continuum #5: Neurofunk drum ʻnʼ bass versus speed garage C27.P128
(1997). The Wire 398. https://www.thewire.co.uk/in-writing/essays/the-wire-300_simon-reynolds-on-the-hardcore-
continuum_5_neurofunk-drum_n_bass-versus-speed

Rietveld, H. C. (1998). This is our house: House music, cultural spaces, and technologies. Ashgate. C27.P129
Google Scholar Google Preview WorldCat COPAC
Rietveld, H. C. (2018). Machine possession: Dancing to repetitive beats. In C. Levaux and O. Julien (eds), Over and over: Exploring C27.P130
repetition in popular music, ch. 5. Bloomsbury Academic. https://doi.org/10.5040/9781501324871.ch-005
Google Scholar Google Preview WorldCat COPAC

Scheuerman, W. E. (2004). Liberal democracy and the social acceleration of time. Johns Hopkins University Press. C27.P131
Google Scholar Google Preview WorldCat COPAC

Schüll, N. D. (2012). Addiction by design: Machine gambling in Las Vegas. Princeton University Press. C27.P132

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
Google Scholar Google Preview WorldCat COPAC

Sharma, S. (2014). In the meantime: Temporality and cultural politics. Duke University Press. C27.P133
Google Scholar Google Preview WorldCat COPAC

Shaviro, S. (2015). No speed limit: Three essays on accelerationism. University of Minnesota Press. C27.P134
Google Scholar Google Preview WorldCat COPAC

Spicer, M. (2004). (Ac)cumulative form in pop-rock music. Twentieth-Century Music 1(1): 29–64. C27.P135
https://doi.org/10.1017/S1478572204000052
Google Scholar WorldCat

St John, G. (2019). At home in the big empty: Burning man and the playa sublime. Journal for the Study of Religion, Nature and C27.P136
Culture 13(3): 286–313. https://doi.org/10.1558/jsrnc.36778
Google Scholar WorldCat

Stanyek, J., and Piekut, B. (2010). Deadness: Technologies of the intermundane. Drama Review 54(1): 14–38. C27.P137
https://doi.org/10.1162/dram.2010.54.1.14
Google Scholar WorldCat

Thomson, I. D. (2011). Heidegger, art, and postmodernity. Cambridge University Press. C27.P138
Google Scholar Google Preview WorldCat COPAC

Thornton, S. (1995). Club cultures: Music, media and subcultural capital. Wiley. C27.P139
Google Scholar Google Preview WorldCat COPAC

Turner, V., Abrahams, R. D., and Harris, A. (2017). The ritual process: Structure and anti-structure. Routledge. C27.P140
https://doi.org/10.4324/9781315134666
Google Scholar Google Preview WorldCat COPAC

Virilio, P. (1991). The aesthetics of disappearance (trans. P. Beitchman). Semiotexte. C27.P141


Google Scholar Google Preview WorldCat COPAC

Wark, M. (2017). Black accelerationism. https://publicseminar.org/2017/01/black-accelerationism/ (accessed 2.22.20). C27.P142


Google Scholar Google Preview WorldCat COPAC

Witek, M. A. (2017). Filling in: Syncopation, pleasure and distributed embodiment in groove. Music Analysis 36(1): 138–160. C27.P143
https://doi.org/10.1111/musa.12082
Google Scholar WorldCat

Witek, M. A. (2019). Feeling at one: Socio-a ective distribution, vibe, and dance-music consciousness. In R. Herbert, D. Clarke, C27.P144
and E. Clarke (eds), Music and Consciousness, vol. 2, 93–112. Oxford University Press.
https://doi.org/10.1093/oso/9780198804352.003.0006

Witek, M. A., Clarke, E. F., Wallentin, M., Kringelbach, M. L., and Vuust, P. (2014). Syncopation, body-movement and pleasure in C27.P145
groove music. PloS one 9(4): e94446. https://doi.org/10.1371/journal.pone.0094446
Google Scholar WorldCat
Wittgenstein, L., and Anscombe, G. E. M. (1953). Philosophical investigations. Macmillan. C27.P146
Google Scholar Google Preview WorldCat COPAC

Žižek, S. (2012). Organs without bodies: On Deleuze and consequences. Routledge. C27.P147
Google Scholar Google Preview WorldCat COPAC

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471673 by National Science & Technology Library user on 26 May 2023
The Oxford Handbook of Time in Music
Mark Do man (ed.) et al.

https://doi.org/10.1093/oxfordhb/9780190947279.001.0001
Published: 2021 Online ISBN: 9780190947309 Print ISBN: 9780190947279

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
END MATTER

Index 
Published: December 2021

Subject: Music
Series: Oxford Handbooks
Collection: Oxford Handbooks Online
Index

“Due to the use of para id indexing, indexed terms that span two pages (e.g., 52–53) may, on occasion,
appear on only one of those pages.”
Note: Tables and gures are indicated by t and f following the page number
À la recherche du temps perdu (Proust) 38–39

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
Ableton Live software 342–43343f
absolute time 58–59254352
accelerationism 19570571576586
Ackermann, William 312–13
action-mirroring 156–57
Aeschylus 2–3
aesthesic, from Tagg 418–19
aesthetics
experience of 86445–46
function of distraction 132–37
of politics 102–3
Aesthetics (Schleiermacher) 34–35
a ective musical time 9193–94
anticipatory systems 174
Apartment House 515–22
Arendt, Hannah 122–23423–24
Aristotle 65330
Ars cantus mensurabilis (‘The Art of Measured Song’) (Cologne) 3n2
Art of the Musician (Hanchette) 319
Ashton, Frederick 416
Attack Magazine 344
attunement 93–95100150–51181–84507580–81
autopoiesis theory 174243
Bach, Johann Sebastian 287–88
Brandenburg Concerto No. 6 (Bach) 13–14287–88
Bacilly, Bénigne de 4
Bailey, Derek 531–32
Ballard, Glen 341357
ballet musical time.
Barad, Karen 9193n4–94
Barnum, P.T. 308
Baudelaire, C. 50
beatmatching 235–47
Becker, Judith 446
Beethoven, Ludwig van
Egmont Overture 28
metronomes and 315n9324n39328–29
Symphony No. 5 in C minor 42
Symphony No. 9 in D minor 42113118122123–24
Wellingtons Sieg 311
Being and time (Heidegger) 528
Benesh Movement Notation (BMN) 419–20419n10–20
Berger, Karol 554
Bergson, Henri 35–36447–48528530–31534541
Berlioz, Hector 42323
Berry, Chuck 366
Berry, Wallace 111
Besseler, Heinrich 35–36
Beyond Structural Listening (Dell’Antonio) 141
Beyond the Pleasure Principle (Freud) 53–54

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
Bizet, Georges 429f429
Black talk (Sidran) 527–28
Blake, Blind 369
blues
country blues 371–73
8-bar blues 374
16-bar blues 373380–81
standard blues forms 373–74
The country blues (Charters) 371–72
12-bar blues 367373375–78379–81384
Blues fell this morning (Oliver) 371–72
Boethius 23
p. 594 Boretz, Benjamin 94
Borgo, David 487
Born, Georgina 447–48
Bowman, E.M. 320–21
Bowman, Wayne 487–88
Brahms, Johannes 43319n25–20
Broonzy, Big Bill 383
Brower, Harriette Moore 321
Brown, James 79f79–8080f82–8386
‘Sex Machine’ (Brown) 79–8080f
‘Soul Power’ (Brown) 79f79
Brown, Willie 369378379
Buchner, Alexander 309
Bülow, Hans von 52
Butler, Judith 92
Cage, John 18127131–32
Concert for Piano and Orchestra (Cage)
Carr, Leroy 369370–71383
carreteiro rhythm 179
Casals, Pablo 106
Central Javanese Classical gamelan. See palaran in Central Javanese Classical gamelan
Charters, Samuel 371–72
Chittenden, Kate S. 319–20
Chopin, Frédéric
Ballade No. 2 in F major 6263f
Étude Op. 25 No. 11 in A minor, ‘Winter,’ 62
Piano Sonata No. 2 in B♭ minor (Chopin) 54
Preludes for piano 742
Preludes for piano (Chopin) 51–5253–54
chronoi (basic units of time) 2–3
chronos 25507–8
chunking 11
Clarke, Arthur C. 53–54136–37
click tracks 245–46326345349–52507–8
clock time 5859312320342509516–17579–81
cognition 135141244
Collins, Sam 373380–82

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
‘Loving lady blues’ (Collins) 381–82
Cologne, Franco de 3n2
Columbia Records 383
Compendium Musicae (Descartes) 26–27
complex adaptive systems theory 174
ComputerRhythm (Eko) 348
concurrit style 3
Cone, Edward 434
Cook, Nicholas 100111–12123–24507529–30531–32
Cox, Ida 369–70
cross-modal correspondences 215215n1221–24227–30
Crumb, George 70f70–71
Idyll for the Misbegotten (Crumb) 70f70–71
Csíkszentmihályi, Mihály 506
Cunningham, Merce 517
cyborg theory 542
cypher notation system 469–70
Dasein, de ned 583–85
Davis, Miles 106
deep time in music. See epochal time
Dela eld, John 308
Deleuze, Gilles
death drive 53–54113120–21
dynamic vs. static repetition 85–86
musical time and 919294n597n1197–101102n19–3
on musical work 114–16120–22
politics of musical time 8–991n197–10198n13102n19–3
DeNora, Tia 461–62
Descartes, René 26–28
Dewey, John 31–3233
Dewsbury, J.D. 422
Dickenson, Emmett 376–78
digital audio workstations (DAWs) 343–44
Dionysius of Halicarnassus 2–3
direct dissonance 390406412
disattunement 181–84
disco 120579–81
discrete musical events 157n12–58216–18
discrete pitch 216–17
displacement in metrical dissonance 284–93286t290t390
dissonance. See metrical dissonance
distraction 111
The Problem of Distraction (North) 130141–42
p. 595 Do man, Mark 507–8
Dowling, Paul 419–20
drum and bass (DnB) 567
Ducret, Marc 491
Dusman, Linda 112
dynamic attending 154155–56236–37246254581

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
dynamic systems theory 10–11155–56244
Earth, Wind & Fire 9116–20
‘September’ (Earth, Wind & Fire) 9116–20
electroencephalography 244–45
electromyographic (EMG) data 206–7211
electronic dance music (EDM) 79280–81568–71574–75577–79586–88
electronic music 354
Elements and Notation of Music (McLaughlin) 320
Elias, Norbert 424
emotional expression 157–58220228–29
empathetic attunement 507
en-temps (in time) 201–2
enactivism theory 235
Entrainment 242–47
enactive coupling 235–36242–47
motor entrainment 245–46
social entrainment 246
Eonta (Xenakis) 135–36
epochal time 741–424647–4849–50
Eshun, Kodwu 354–55579
event time 98171
everyday time 277549552568
expectation and music 4–5
experience of time
Bergson on 533
dynamic attending 154155–56
evolution of music and 150–51
introduction to 149150–51
lyric time and 64–65
musical motivation and 158–59
musical pulse and 154–57
musical time and 159–60
musical tone and 157–58
re ections on 152–54
relationship to music 13–47–8
sense of movement and 539
subjectiveness of 67–6868n1769n1869–7071
technological acceleration and 571–72
Western modernity impact on 567
Extended Mind Theory 245
Fahey, John 379
falsobordone style 3
Fantasia in C major Hob XVII4 Op. 58 (Haydn) 129
Faure, Sylvia 421–22
Feldman, Morton 19550–51554–63558f
String Quartet No. 2 (Feldman) 550–51554–63558f
Fink, Robert 569
Fontana, Bill 19550–53560–63
Foucault, Michel 561

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
‘Four-on-the- oor’ rhythm 342–43343f
Franck, César 74649–50
Freud, Sigmund 53–54120–21130–31
Fuller, Blind Boy 373
funk 886
Galin-Paris-Chevé notation system 469–70
Gallope, Michael 114115–16
game theory 582
The Gay Science (Nietzsche) 33
Genetic Choir 529538–40
German metre 427–28432
gestures and music 150–51
Goerner, Nelson 139–40
Gould, Hanna F. 309–10
Grandmaster Flash 235
‘Great’ Symphony No. 9 in C major, D 944 (Schubert) 28
Green, Lucy 422–23
Greenberg, Clement 557
Gress, Drew 487489–501
groove
the ‘changing same’ in 8
computer-based grooves 82
intensi cation of 85–86
listener engagement in 79
state of being-in-the-groove 8586–8788
grouping dissonance 280f280–81281f296–98297t
habit, de ned 185185n21
Hamelin, Marc-André 129
Hamilton, C. G. 308–9311
p. 596 Hammond, John 383
Hanchette, Henry G. 319
Handy, W.C. 366–67369–70384
Harmonia, de ned 25–26
‘Harmonic Pendulum or Time-Gage’ (Light) 313
harmonic rhythm 101391393396–97402405–6410–12430–31
Harris, Calvin 280–81281f
‘I’m Not Alone’ (Harris) 280–81281f
Hasty, Christopher 434569
Halbwachs, Maurice 527
Haydn, Joseph 48129
London Symphonies (Haydn) 48
Heidegger, Martin 353777120–21528583–85
Helias, Mark 487
Henze, Hans Werner 416
Higgins, Billy 389392–96398–401
Highmore, Ben 550
Hill, King Solomon 369
Hindemith, Paul 201
historical time 716–17120170177f180445–46449453456–57459552

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
Hodges, Nicolas 135–36
Hø ding, Simon 507
Holcroft, Thomas 310
Holland, Bernhard 324–25
homorhythm 393
hors-temps (outside time) 201–2
House, Son 369375378
Howlin’ Wolf 378379
The Human Condition (Arendt) 423–24
human movement science 197–98
human phylogenetic timescale 170
Husserl, Edmund 35130–31132–33137–38197202
Hutton, James 46
Hynd, Ronald 416
improvisation 15488
indeterminacy
‘In Memoriam’ (Tennyson) 47
Indonesian gamelan (palaran) See palaran in Central Javanese Classical gamelan
information-processing theory 155–56242–43
Instant Composers Pool (ICP) Orchestra 529531–32534–36
inter-onset-intervals (IOI) 224–25
Ishiguro, Kazuo 127–28142–43
isochrony in rhythm and metre 254–57256f
Jackson, Papa Charlie 370–71
James, Skip 375
Jameson, Fredric 549
jazz
‘changing same’ in 8
cognitive processes of performance 488
communication with audience 496–97498
improvisation in 8488
works of 116–20390
jazz musicians 17–18478–79497540543–44
Je erson, Blind Lemon 369372–73374–78376t377t
Matchbox blues’ (Je erson) 375–78376t377t
Johannsen, Thomas 529538–40
Johnson, Robert 367368369378382–83384
Johnson, Tommy 369
Jones, Mari Reiss 581
Jordan, Louis 366
Jordan, Stephanie 433
jungle 571–73
kairos, de ned 25507–8
Kant, Immanuel 28–323437130–32
Karajan, Herbert von 305
Kau man, Stuart 173
Keightley, Emily 447–48
Kempelen, Wolfgang von 308–9
Kirnberger, Johann Philipp 351–52

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
Kobranie system 529536–38
Kone, Drise 257–58
Kotík, Petr 513–14
Kronos Quartet 554–55562
Kuyate, Jeli Madi 257–59
Langenus, Gustave 320
Large, Edward 356–57
L’Arlésienne suite (Bizet) 429f429
Latour, Bruno 541
Leadbelly 371376–78
Lefebvre, Henri 550553–54560–61
Light, Thomas 313
Linn, Roger 345–48346f
p. 597 listeners/listening.
aesthetic experience of 445–46
intentionality in 199
musical instants and 198–200
musicological vs. musical 418–19
psychology of 13–14
temporality of 139
trivialised listening 136
Liszt, Franz 43134–35
Tre Sonetti del Petrarca S.161 (Liszt) 134–35
LM-1 Drum Computer (Linn) 345–49346f352
Locke, Alain 367–68
Logos, de ned 25–26
Lomax, Alan 371
Lomax, John 371
Lorimer, Hayden 422
Lyell, Charles 46–47
lyric time 7–84261–65
macapat timing 468–69470–75479481–82
macatan singing tradition 17
Machine Age 321
macro timescales 201202
Maelzel’s Automaton Turk 309–11310f
Maelzel’s Panharmonicon 311–12311n5318
magnetoencephalography 244–45
Mahler, Gustav 4849
Symphony No. 8 in E♭ major (Mahler) 48
Malaby, Tom 485486487
Malian music 266–69
Mankekar, Purnima 460
Manovich, Lev 349
Marx, A.B. 326
Massumi, Brian 95
McCormack, Derek 422
McDowell, Fred 375
McKune, James 369

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
McLaughlin, James Matthew 320
McLuhan, Marshall 527528
melodic contour 198468471473476–77480481
Mendelssohn, Felix 42–43
Mengelberg, Misha 531–32534–36
mensural system 3–466–67351–52
Merleau-Ponty, Maurice 35–38132–33139
meso timescales 201202207
Messiaen, Olivier 5765–6767f70–71
‘Louange à l’éternité de Jésus’ (Messiaen) 66–6767f70–71
metre/metrics
absence of 66
analysis in ‘Evidence,’ 405–12407f408f409f410f411f
binary metre 278–79
Franco-Italian hypermetre 427–28432433–34
German metre 427–28432
isochrony in rhythm and metre 104–5254–57256f
mixed metres 287293–96294f295t
pucung metre 471
metric accentuation 319
metric displacement. See displacement in metrical dissonance
metric theory 256–57
metrical dissonance
metronomes.
adoption of 319–22
auto-metronomic rhythm and 312–19
clockwork metronome 303306307–8313314–17320–21330f330350–51
in groove context 82–83
tape-measuring ‘portable metronome,’ 313
Metzger, Heinz-Klaus 555
microrhythms 33n3255
microtiming 104–5155–56253255390
MIDI sequencer 342–43343f
milliseconds 80–81169–70171304–5343–44
mind wandering 141
mindfulness 141
Mingus, Charles 492
‘Modern Piano Methods’ (Elson) 320
Molecular bonding timescale 171
Monk, Thelonious.
Evidence (Monk) 389–414
I’m Confessin’ (Monk) 390
Misterioso’ (Monk) 391
morphodynamical theory 208
Moscheles, Ignaz 316
motor/motion control 197–98199203–4206–7
‘motor units’ of time 324–25
p. 598 motormimetic cognition 199
MPC60 (Akai) 356

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
Mullov-Abbado, Misha 116–20
musical instants 200
The Musical Magazine 317
Musser, Clair Omar 326
Nash, Catherine 422–23
natural law/natural rhythm 2–331–32
neoliberalism 549–50
neuropsychological studies 154–55
New York performance network 487–501
Newtonian concept of time 58–59
Nichomachean Ethics (Aristotle) 330
Nietzsche, Friedrich 333648–49
noema of sound 133136138
non-representational theory (NRT) 16422–24
North Indian classical music 446–59
novelistic time 743–50
Nussbaum, Martha 37–38
The Nutcracker (Tchaikovsky) 425–29426f427f
Obbligato quintet 498–500499f
Oliver, Paul 371–72
Ondine (Henze) 416
Ong, Walter 528
ontogeny, de ned 187
Open Loose trio 500–1
Ore, John 392
Oscillating Steel Grids along the Brooklyn Bridge (Fontana) 19550–53560–63
overtone series 277–78
‘P-Funk (Wants to Get Funked Up)’ (Parliament) 80f80–8381f
Paderewski, Ignacy Jan 318–19326
PAiA Programmable Drum Set 348
palaran in Central Javanese Classical gamelan 467–78
Panharmonicon (Maelzel) 311–12311n5318
Parchman Farm State Penitentiary 371
Parsons, A.R. 316–17
participatory systems 38–39181–82353–54
Patton, Charley 367378–80380t381t
‘Moon going down’ (Patton) 379–80380t
Penrose, Roger 172
perception, de ned 243–44
Perform or Else: From Discipline to Performance (McKenzie) 131
autonomous performer behaviour 513–15
of bodies 929394
cognitive processes in jazz 488
contemporaneous performance practice 3n1
cues in 139–40
distraction and 130131–32141–42
music/performance/social event timescale 171
musical time 139194104
raw expression of live performances 355

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
stile rappresentativo (stile parlando) in 3
timings of ensemble drumming 264–65
timings of jembe drumming 263–64
unconsoled performers 127–30
‘perfumance,’ 131–32
Pettengill, Richard 112
phase correction 155–57238–39241–42244–46
phase-locking 222244247
phase transitions 204–5205f
Philosophy of Art (Schelling) 29
The Philosophy of Manufactures (Ure) 318
pitch
contours of 209
discrete pitch 216–17
high pitch 200219–20374
kempul pitches 474475–81
kenong pitches 474475–81
Planck timescale 171174
Platonic idealism 114115
Platonist conceptions of music 225–2629–3037
PLUR (peace, love, unity, and respect) scene 572n7577
Poe, Edgar Allen 309
‘Poetic license’ (Gress) 495–96496f
poiesis, de ned 102–3
poïetic, from Tagg 418–19
Pouillaude, Frédéric 423–24
prelude-machine 52
Principles of Geology (Lyell) 46
The Problem of Distraction (North) 130141–42
p. 599 progressive time 7–861–65
proto-phenomenological framework of time 534
psychoacoustics research 344
pucung metre 471
Pythagorean conceptions of music 25–26
quantization
Quatuor pour la n du temps (Messiaen) 66–6767f70–71
Questlove (drummer) 356
‘Race Records’ industry 369–70
Radiolab 327–29
Rainey, Ma 369–70
Rainey, Tom 485–86487–501
Ram, Kalpana 450
Rancière, Jacques 84491n195–96101–6102n19–3
rasikas (connoisseurs) of listening 448–61
Régulier, Catherine 553–54560–61
Renaissance 25–26
Resident Advisor 344
Revill, George 422–23
rhythm

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
basic research into 77–78
carreteiro rhythm 179
categorical rhythm perception 12–13
de ned 77
‘Four-on-the- oor’ rhythm 342–43343f
harmonic rhythm 101391393396–97402405–6410–12430–31
homorhythm 393
hyper-rhythm 579
isochrony in 254–57256f
microrhythms 33n3255
natural law/natural rhythm 2–331–32
polyrhythm 236577579582
Rhythm and Tempo (Sachs) 324
Rhythm-Builder (Langenus) 320
rhythmic quantization See quantization
Rhythmanalysis (Lefebvre) 422
rhythmic displacement 401–2404–5409412–13
rhythmopoetic in ection 319
rhythmos, de ned 25–2683
Ricoeur, P. 84
Rijke, Ralph de 538–39
Rink, John 111112
ritornello form movement 288289f
Roberts, Paul 134–35
Rock, Bob 355
Roman Empire 2
Romantic era 61–62314
Rouse, Charlie 392401–5
Rousseau, Jean-Jacques 27–28
Royal Academy of Dance (RAD) 419–20
Sachs, Curt 324
Sawyer, Keith 506–7
Schae er, Pierre 199
Schelling, F.W.J. 29–31333437–38
Schindler, Anton 317324n39
Schlegel, Friedrich 32–33
Schleiermacher, F.D.E. 34–35
Schole eld, Daniel 313
Schubert, Franz 424449–50
‘Great’ Symphony No. 9 in C major, D 944 (Schubert) 28
Symphony No. 9 in C major (Schubert) 74445
Schumann, Clara 98–10099f
Schumann, Robert 42–4344–4549–5062–63
Symphony No. 4 (Schumann) 45
Symphony No. 9 in D minor (Schumann) 7
Three Romances for Violin and Piano (No. 1) Op. 22 (Schumann) 98–10099f
‘Träumerei’ from Kinderszenen Op. 15 (Schumann) 62–63
Schütz, Alfred 507–8529–33540–44
‘Making Music Together’ (Schütz) 529–33

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
scratching in DJing 239n4
A Sea Symphony (Williams) 48
Seashore, Carl Emil 324–25
Seddon, Frederick 507
Self-Help in Piano Study (Brower) 321
7 Black Butter ies (Gress) 489–96
Shepherd, John 422–23
shu e rhythm 253259–61260f
Sidran, Ben 527–28
Simondon, Gilbert 344–45
Simonen, Sini 105
skeuomorphical rendering of virtual piano roll 342
p. 600 slip-cueing 239–42
Smith, Bessie 369–70
Smith, Harry 369
Smith, Zadie 416418421
Swing Time (Smith) 416418421
Songs without Words (Mendelssohn) 42–43
sonic heterochronies 560–62
sound
accompanying body motion 200
constraints 203
noema of 133136138
producing body motion 203–4
propagation of sound constraints 203
sustained sound 204
understandings of time and 446–47
wah-wah sound 210
spans, de ned 10–11
speech act theory 84–85
speech perception 268–69
Spicer, Mark 569
Spohr, Louis 316–17
Grand Violin School (Spohr) 316–17
Sterne, Jonathan 528
stile rappresentativo (stile parlando) in performance 3
Strong, George Templeton 311–12
subjective time 58–5959t84152–53
Sulcas, Rosalyn 421
‘The Swan’ (Baudelaire) 50
Swan Lake (Tchaikovsky) 417f417423–24
The Swan (Saint-Saëns) 216f216
Symphony in D minor (Franck) 74649–50
synchronization 2–3237–38313See also entrainment
sensorimotor synchronization 155156–57235–36238–40246–47255
syncopation 6680–81117211236280477574
synthesis of time 94–9597–101104106121–22
systematic variation 85–86
Taborn, Craig 491492–93

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
Takt, de ned 36
Tb-303 synthesizer (Roland) 569–70
technomorphism 351
teleodynamics 174
tempo-dependent system for pulse salience weightings 291–92291t292f
tempo giusto (‘just’ or ‘suitable’ time) 3–4
temporal automata 14
temporal ‘consonance,’ 2–3
temporal correction in beatmatching 240–42
temporal di erentiation 33–34587–88
temporal extension 188197–98
temporal indeterminacy 18505–6512519521
temporal jouissance 582–84
temporality/time in music
absolute time 58–59254352
clock time 5859312320342509516–17579–81
consciousness of 87153
de ned 149
epochal time 741–424647–4849–50
event time 98171
everyday time 277549552568
historical time 716–17120170177f180445–46449453456–57459552
human time 365–66555
linear timekeeping 155–56556–57
lyric time 7–84261–65
macapat timing 468–69470–75479481–82
of meaning 87
microtiming 104–5155–56253255390
milliseconds 80–81169–70171304–5343–44
novelistic time 743–50
objective time 58–5959t
subjective time 58–5959t84152–53
temporal organization 12–16446
‘Ten Theses on Politics’ (Rancière) 95–96
Tennyson, Alfred 47
Thanatos and repetition 121
Thayer, Alexander Wheelock 308311n5
thesis-arsis motion 351–52
Thrift, Nigel 422
Thus Spoke Zarathustra (Nietzsche) 48–49
tiered timescales 170–72
time. See temporality/time in music
Time and Free Will (Bergson) 528
p. 601 timescales
constraints in 198202–5
macro timescales 201202
mapping timescales and emergence 170–76
meso timescales 201202207
micro timescales 201

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
music/performance/social event timescale 171
musical instants and 200–2
reductionism critique 169172–73
summary of 188–89
understanding of duration through 10–11
Universe timescale 170174
TR-808 (Roland) 348
Transcranial Magnetic Stimulation (TMS) 226
Transient neurodynamic timescale 171
12-bar blues 367373375–78379–81384 See also blues
The Unconsoled (Ishiguro) 127–28142–43
United Kingdom’s Musicians’ Union 354
Urban, Greg 490
Ure, Andrew 318
Van Ditzhuyzen, Martine 539–40
Varela, Francisco 487
Verdi, Giuseppe 428430f430–34431f
Otello (Verdi) 430f430–34
Rigoletto (Verdi) 428
Village Voice 354
virado movement during drumming 179
virtuosity 412575–76
Wagner, Richard 4249
Parsifal (Wagner) 49
wah-wah sound 210
Waller, Fats 366
Waters, Muddy 378
Ways of Listening (Clarke) 136–37
Weber, Carl Maria von 316
Weber, Gottfried 314314n8326
Webster, Daniel 318
‘Weise menschlichen Daseins’ (‘manner of human existence’) 35–36
Whelan, Pete 369
White, Bukka 375
Whitman, Walt 48
Williams, Vaughn 4849
Willis, Richard Storrs 312
wipe-out phenomenon 418
Wirt, William 308
Withers, Bill 279f280
‘Lean on Me’ (Withers) 279f280
Wordsworth, William 43
working memory 135150–51
WS-1 write switch (Roland) 348
Xenakis, Iannis 135–36201–2
‘Zaftig’ (Gress) 492–94493f
Zapolska, Tanya 105
Zeisler, Fannie Bloom eld 318–19
Zižek, Slavoj 580

Downloaded from https://academic.oup.com/edited-volume/41628/chapter/353471804 by National Science & Technology Library user on 26 May 2023
zoē, de ned 122–23
zukommen concept 583

You might also like