Film Music: Cognition to Interpretation [1 ed.] 1138586706, 9781138586703

Film Music: Cognition to Interpretation explores the dynamic counterpoint between a film’s soundtrack, its visuals and n

135 102 27MB

English Pages 252 [261] Year 2024

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Film Music: Cognition to Interpretation [1 ed.]
 1138586706, 9781138586703

Table of contents :
Cover
Half Title
Title
Copyright
Contents
Acknowledgments
Introduction
1 Empathy
2 CONTAINER Schema
3 LINEARITY Schema
4 SOURCE-PATH-GOAL & CONTAINER Schemas
5 Affordances
6 Memory & Auditory Perception
7 Archetypes
8 Associations
9 Categorization
10 Interpretive Transformations
Appendix I Empathy
Appendix II Conceptual Metaphor and Image Schema
Appendix III Affordances
Appendix IV Memory
Appendix V Auditory Perception
Appendix VI Archetypes
Appendix VII Conceptual Integration
Appendix VIII Categorization
Appendix IX Additional Examples
References
Film, Filmmaker, and Composer Index
Subject Index

Citation preview

FILM MUSIC

Film Music: Cognition to Interpretation explores the dynamic counterpoint between a film’s soundtrack, its visuals and narrative, and the audience’s perception and construction of meaning. Adopting a holistic approach covering both the humanities and the sciences—blending cognitive psychology, musical analysis, behavioral neuroscience, semiotics, linguistics, and other related fields—the author examines the perceptual and cognitive processes that elicit musical meaning in film and breathe life into our cinematic experiences. A clear and engaging writing style distills complex concepts, theories, and analytical methodologies into explanations accessible to readers from diverse disciplinary backgrounds, making it an indispensable companion for scholars and students of music, film studies, and cognition. Across ten chapters, extensive appendices, and hundreds of film references, Film Music: Cognition to Interpretation offers a new mode of analysis, inviting readers to unlock a deeper understanding of the expressive power of film music. Juan Chattah is Associate Professor of Music at the Frost School of Music, University of Miami, USA.

FILM MUSIC Cognition to Interpretation

Juan Chattah

Designed cover image: © Juan Chattah First published 2024 by Routledge 605 Third Avenue, New York, NY 10158 and by Routledge 4 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN Routledge is an imprint of the Taylor & Francis Group, an informa business © 2024 Juan Chattah The right of Juan Chattah to be identified as author of this work has been asserted in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988. All rights reserved. No part of this book may be reprinted or reproduced or utilised in any form or by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying and recording, or in any information storage or retrieval system, without permission in writing from the publishers. Trademark notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe. Library of Congress Cataloging-in-Publication Data Names: Chattah, Juan, author. Title: Film music: cognition to interpretation/Juan Chattah. Description: [1.] | New York: Routledge, 2023. | Includes bibliographical references and index. Identifiers: LCCN 2023029274 (print) | LCCN 2023029275 (ebook) | ISBN 9781138586703 (hardback) | ISBN 9781138586710 (paperback) | ISBN 9780429504457 (ebook) Subjects: LCSH: Motion picture music—Analysis, appreciation. | Motion picture music— Psychological aspects. Classification: LCC ML2075. C468 2023 (print) | LCC ML2075 (ebook) | DDC 781.5/42—dc23/eng/20230621 LC record available at https://lccn.loc.gov/2023029274 LC ebook record available at https://lccn.loc.gov/2023029275 ISBN: 978-1-138-58670-3 (hbk) ISBN: 978-1-138-58671-0 (pbk) ISBN: 978-0-429-50445-7 (ebk) DOI: 10.4324/9780429504457 Typeset in Optima by Apex CoVantage, LLC

CONTENTS

Acknowledgments

vii

Introduction

1

1 Empathy

6

2 CONTAINER Schema

23

3 LINEARITY Schema

37

4 SOURCE-PATH-GOAL & CONTAINER Schemas

50

5 Affordances

64

6 Memory & Auditory Perception

85

7 Archetypes

100

8 Associations

116

9 Categorization

125

10 Interpretive Transformations

146

vi

Contents

Appendix I

Empathy

165

Appendix II

Conceptual Metaphor and Image Schema

172

Appendix III

Affordances

181

Appendix IV

Memory

187

Appendix V

Auditory Perception

194

Appendix VI

Archetypes

203

Appendix VII

Conceptual Integration

212

Appendix VIII Categorization

217

Appendix IX

223

Additional Examples

References

229

Film, Filmmaker, and Composer Index

245

Subject Index

248

ACKNOWLEDGMENTS

Behind the scenes of crafting this work, a cast of extraordinary individuals and institutions graciously extended invaluable support, guidance, and expertise, transforming the writing process into an awe-inspiring journey. April Mann, your boundless dedication to the craft of writing guided me through the labyrinth of words, your discerning questions became sparks that ignited a kaleidoscope of ideas, and your sincere friendship enriched this journey in ways that words cannot fully express. I am deeply indebted to a vibrant community of colleagues and friends working in film music and cognition, as their inspiring work and invaluable feedback stimulated illuminating conversations and opened new avenues of exploration. My heartfelt thanks go to Josh Albrecht, Mihailo Antovic´, Michael Austin, William Ayers, José Luis Besada, Daniel Bishop, Janet Bourne, Yorgason Brent, Warren Buckland, James Buhler, Poundie Burstein, Ewan Alexander Clark, David Clem, Maarten Coëgnarts, John Covach, Jeanine Cowen, Mark Cross, James Deaville, Janice Dickensheets, Kevin Donnelly, Rebecca Eaton, Zohar Eitan, Robert Fink, Lucio Godoy, Julianne Grasso, Michele Guerra, Erik Heine, Guido Heldt, Hubert Ho, Julie Hubbert, Bryn Hughes, Deniz Hughes, David Huron, Jesse Kinne, Patrick Kirst, Violetta Kostka, Mariusz Kozak, Peter Kravanja, Danijela Kulezic-Wilson, Gui Hwan Lee, Frank Lehman, Scott Lipscomb, Charity Lofthouse, Justin London, Catherine Losada, Sean McMahon, Kate McQuiston, Miguel Mera, Táhirih Motazedian, Scott Murphy, Chelsea Oden, Mitch Ohriner, Steven Rahn, Ron Sadoff, Janna Saslaw, Tom Schneller, Matthew Shaftel, Dan Shanahan, David Shire, Robynn Stilwell, Siu-Lan Tan, Rennee Timmers, Yayoi Uno Everett, Petros Vouvaris, Elsie Walker, Emile Wennekes, and Lawrence Zbikowski. In the fervor of expressing gratitude, I may have unintentionally left out individuals who have contributed significantly to bringing this book to fruition. To those inadvertently omitted, please know that your support, guidance, and contribution are deeply appreciated. I am indebted to the University of Miami, in particular to Charles Mason, Chair of Music Theory and Composition, Shelly Berg, Dean of the Frost School of Music, and Guillermo (Willy) Prado, Executive Vice President for Academic Affairs and Provost. Their steadfast support and commitment to fostering intellectual growth allowed me to expand my horizons and develop ideas beyond the confines of traditional disciplines.

viii

Acknowledgments

On a personal note, I am deeply grateful to my mom, ‘La Chely’, whose companionship throughout the writing process has been a source of immense comfort and motivation. Despite the distance, especially during those long and challenging COVID years, your virtual presence (and thrilling updates on ‘la novela’) made this solitary journey much more enjoyable. Israel, you are my life compass. I am eternally grateful for your unwavering love and grounding presence, which have been the guiding stars that kept me on course throughout this writing odyssey, and for your selfless and endless patience, which anchored me amidst the fluctuations of inspiration and gave me the stability and balance needed to see this book through to its completion. As I conclude these acknowledgments, I want to extend my deep appreciation to Routledge and its extraordinary ensemble of professionals—particularly Helen Birchall and Autumn Spalding—for making this book a reality, and last but not least to you, the reader, who I hope will be inspired to delve into the depths of knowledge and embrace the infinite possibilities that lie within the realms of film, music, and the fascinating wonders of the human mind.

INTRODUCTION

At long last! The opening night of Star Wars: Episode VI – Return of the Jedi. The lights inside the movie theater begin to dim. A majestic fanfare accompanying the emblematic yellow rollup of narrative crawling into the infinite darkness of space invites us into the film’s fantastical realm. The camera pans to reveal an Imperial Star Destroyer moving toward an intimidatingly massive Death Star. As a smaller shuttle hurtles from the Destroyer toward the Star, the captain instructs, “Command station: This is ST 321. Code clearance: Blue. We’re starting our approach. Deactivate the security shield.” In the music, a powerful brass ensemble announces, “BOM BOM BOM, BUM BI-BOM”, accents of impending menace and wrath, repeated lower and lower. At the Star, a commander rushes through a formation of stormtroopers over to the gate, uneasy, anticipating the shuttle’s arrival. The shuttle lands. As the shuttle’s hatch whooshes open, revealing its inner darkness, the music strikes its lowest, most menacing notes, extending and completing the initial statement, “BOM BOM BOM, BUM BI-BOM, BUM BI-BOM”. The commander shivers, turning pale as he hears heavy footsteps and mechanical breathing emerging from the depths of the shuttle—Darth Vader, Lord of the Sith, has arrived. From the opening scene, the cinematography immerses us in a unique realm wherein our psychophysiological responses subliminally circumscribe our interpretations. The music is vital in shaping these interpretations: through its melody, we intuit Darth Vader’s presence long before he appears on-screen; through its meter, we anticipate an encounter of martial rivalry; through its register and loudness, we sense Darth Vader’s ominous and overpowering temperament and resonate with other characters’ fears, feeling their anxiety firsthand. While the music affects us subliminally, below the level of conscious attention, we sometimes externalize our responses, grabbing a friend’s arm or gasping overtly. Most surprisingly, despite the seemingly negative valence of our responses to the Star Wars scene discussed here, we genuinely enjoy the sinisterly thrilling experience. This remarkable phenomenon is not unique to a single film or scene. In all films, music operates almost subliminally, activating sensorimotor reflexes and conveying complex messages that motivate, support, highlight, complement, or even negate other facets of the cinematic experience. Through the music, we suffer the protagonists’ pain, share their laughter,

DOI: 10.4324/9780429504457-1

2

Introduction

shriek along with their fear, and bond with them through their cinematic journeys. We all recognize that film music is designed to elicit a particular response, but how exactly does the music influence our interpretations of scenes or entire films? What is the source of the music’s expressive power? Are we hardwired to respond to it? To what extent do cultural practices and social norms govern our responses? To answer these questions, we must trace the logic underlying the perceptual and cognitive processes that elicit musical meaning in film.1 This book’s primary goal is to construct a comprehensive framework to explore those processes. Constructing such an ambitious framework requires a holistic approach, one that navigates through the humanities and sciences alike, one that draws on converging thoughts and disciplinary lines—music analysis, psychology, behavioral neuroscience, semiotics, cognitive linguistics, and related fields— foregrounding their implications for understanding the mechanisms whereby film music contributes to our ability to construct meaning.2 I dub this framework ESMAMAPA—Empathy, Schema, Metaphor, Affordance, Memory, Archetype, Personal Association. ESMAMAPA is ambitious in its scope, attempting to explain how interpretations emerge in the listeners’ minds (and bodies) by accounting for the primary cognitive and semiotic processes related to music-based meaning construction. Figure i.1 presents these processes as a taxonomy broadly structured on the relationship between the music and how it generates meaning. While these mechanisms intersect and interact when we construct interpretations, in this book, I consider them one at a time, systematically navigating through this taxonomy, disentangling the complexities of these multiply interrelating mechanisms. Following this introductory chapter, the book echoes the taxonomy presented in Figure i.1, proceeding from (primarily) phenomenological to (primarily) intellectual perspectives. Individual chapters present film vignettes that exemplify and probe the limits of the ESMAMAPA framework, foraying into the extensions and implications of the various mechanisms of meaning construction. Chapters 1 through 5 draw on sub-disciplinary strands of embodied cognition and neuropsychology to explore mechanisms based on analogy, discussing how material characteristics of the music govern (or demand) a specific response from a listener and investigating

FIGURE I.1

Cognitive and semiotic mechanisms underpinning film-music-based meaning construction.

Introduction

3

interactions between the music’s structure and the film’s visual or narrative dimensions. Chapter 1 introduces a musical empathy model that builds on the discovery of the Mirror Neuron System (MNS). Although this phenomenon was first identified in macaques, MNS research reveals neurophysiological mechanisms that might underpin social cognition in humans; and while findings were originally tied to visual stimuli, I draw here an analogy to empathy triggered by aural stimuli. To distill the intricacies of music-based empathy, I further trace our embodied responses to three submechanisms: entrainment, subvocalization, and contagion. Chapters 2, 3, and 4 form a unit delving into Image Schema Theory and George Lakoff and Mark Johnson’s Conceptual Metaphor Theory. These chapters elucidate how film music acts as one agent in a multidimensional mapping process involving the visuals and the narrative. Each chapter centers on a different image schema—CONTAINER, LINEARITY, and SOURCE-PATH-GOAL. Chapter 5 takes James Gibson’s notion of ‘affordances’ as a point of departure, examining musical parameters (meter and tonal centricity) not as properties of the music itself but as emerging in the embodied conceptualization of musical experiences— experiences in which the nature of our bodies determines both what and how music can be meaningful. While Chapters 1 through 5 engage with the phenomenology of music and investigate interactions between the music’s structure and the film’s visual or narrative dimensions, Chapters 6 through 8 begin to extricate the network of associations that emerge from the music, framed by both the audience’s prior experiences and the film’s narrative. These chapters continue to draw on cognitive psychology, but explanations dovetail with theoretical models from cognitive linguistics and speculative notions from semiotics. Chapter 6 discusses the impact of gestalt laws of perception on our recognition and memorization of leitmotifs, drawing a parallel with Pavlov’s classical conditioning and situating leitmotifs as ecologically relevant acoustic cues. Chapter 7 draws on the musicsemiotic theory of ‘musical topics’ and on Charles Sanders Peirce’s and Ferdinand de Saussure’s sign typologies to clarify the nature and function of film music archetypes. Because these archetypes exist outside a single film (inter-opus), becoming cultural units the listener identifies via the music’s stylistic characteristics, this chapter investigates musical topics from both a synchronic and a diachronic perspective. Chapter 8 borrows and adapts the Conceptual Integration model developed by Gilles Fauconnier and Mark Turner to reveal and help reconstruct the thought processes that induce hermeneutic readings of scenes in which the music carries prominent cultural baggage or in which the lyrics explicitly (or implicitly) contain clues for understanding the film. Chapters 9 and 10 expand on the ESMAMAPA framework by exploring two supplementary mechanisms: categorization and transformation. Chapter 9, inspired by Eleanor Rosch’s cognitive models of categorization, disentangles the notion of musical irony and concepts commonly linked to irony, including parody, satire, sarcasm, paradox, hyperbole, and the grotesque. Chapter 10 resumes the discussions on leitmotifs, topics, and conceptual blends, identifying instances where, once established, these undergo transformations that signal new dramatic conditions or plot developments. As a prelude, a film example at the beginning of each chapter sets the space for the mechanisms to be explored. Subsequent examples within each chapter exemplify the mechanism at hand and probe its boundaries. Although these examples represent a wide diversity of approaches (in their musical soundtrack and artistic direction) and a broad range of genres (from silent films to present-day productions), my central goal in choosing them has been to

4

Introduction

illustrate the various cognitive and interpretative mechanisms of the ESMAMAPA framework with utmost clarity. Given music’s shifting role(s) throughout a film, with multiple mechanisms overlapping and interacting to construct meaning, some chapters, particularly later ones, include holistic analyses integrating the theoretical paradigms and critical insights from preceding chapters. Therefore, while advancing through the book and exploring different meaning construction strategies applicable to film music, readers will increasingly find themselves on familiar ground. A fundamental presupposition underlying ESMAMAPA is that humankind’s cognitive capacities and responses to music’s expressive forces are relatively uniform and universal— but the interpretations that we construct stemming from such responses to the music are not. Therefore, when unpacking the film vignettes interspersed throughout the chapters, I do not claim interpretational certainty. Just as my interpretations inevitably reflect my own bodily and intellectual engagement with the music, there are innumerable ways in which audience members may traverse the expressive trajectories purported by the musical underscoring. Furthermore, some of my interpretations, mainly those in later chapters, assume an audience with a certain familiarity with Western musical codes—a familiarity necessary, for example, to identify Darth Vader’s signature melody. Nevertheless, while the creative interpretative act that listeners bring to a film is inevitably informed by their own subjective (personal) and intersubjective (shared) projections and inferences, the ESMAMAPA framework I develop here can be attuned to investigate alternative interpretations. Exploring the mechanisms involved in constructing meaning from film music requires unpacking intricate and sophisticated scholarship. To make this scholarship accessible, I tailor my writing style to a broad audience, using clear prose that includes only the necessary technical language to ensure the depth of treatment and theoretical rigor expected by practitioners and scholars. In terms of music and music-theoretical knowledge, I only assume a basic familiarity with notation and theory fundamentals; this will allow readers to follow along with my transcriptions of selected musical cues. In keeping with the book’s broader goals, I forgo the temptation to delve all too deeply into any theory or cognitive mechanism within the main chapters. Instead, I invite readers to consult the appendices, which offer a more in-depth treatment of scholarship and furnish references to relevant models, methods, and procedures in neuropsychology, cognitive psychology, and semiotics. Nevertheless, even in the appendices, I synthesize these insights with broadly accessible language and examples, optimistic that such prose will also appeal to those with even a passing interest in this scholarship.3 As I completed the manuscript, colleagues teaching film music courses expressed an interest in assigning this book (or selected chapters) as course materials. Their suggestions prompted me to make two additions. First, although all film music can be explored through the ESMAMAPA framework, finding suitable and clear examples for students is no small feat; therefore, I include one appendix featuring a wealth of examples paired with questions and discussion points. Second, although the prose may seem sufficient to grasp the intricacies and subtleties of the arguments put forth in the various chapters, it is no replacement for experiencing the film examples firsthand. Therefore, I invite readers to explore the online by visiting the book’s YouTube site media for all examples that include the play icon accessible via the QR code on Figure I.2 or at www.youtube.com/channel/UCzNQHM7ysgHrkrB4XYyqbA. This dedicated online portal also contains additional examples and expanded explorations of film music.

Introduction

FIGURE 1.2

5

Scannable QR code to access the book’s dedicated online portal.

Notes 1. Readers familiar with my earlier work will recognize traces of ideas and thoughts, yet all proposed frameworks and examples illustrating these frameworks have been significantly revised, refined, and expanded. Therefore, my earlier work offers a prismatic spectrum of concepts and analytical perspectives, from my initial impulse to understand the mechanisms of meaning construction in film music, crystallizing in the proposed framework. 2. In seeking to construct a viable and useful framework, I draw on different kinds of inquiry, each yielding different kinds of evidence, including introspective accounts, composers’ insights, theoretical speculation, and experimental results. I hope that each of these lines of inquiry and evidence will carry its own significance so that readers who do not resonate with my interpretations of a scene, for instance, might nonetheless find merit in the proposed framework or my application of theoretical and empirical scholarship. 3. Much experimental research on metaphor and image schema stems from cognitive psychology or neuropsychology. While research in cognitive psychology investigates behaviors, mental processes, or psychological states, research in cognitive neuropsychology focuses on the corresponding neural substrates. And while cognitive psychology draws on behavioral paradigms (such as reaction time, priming, motion imagery, preferential looking, forced-choice matching, goodness-of-fit) to arrive at results or test hypotheses, cognitive neuropsychology draws on computational modeling and techniques such as functional magnetic resonance imaging (fMRI), electroencephalography (EEG), magnetoencephalography (MEG), positron emission tomography (PET), and event-related potentials (ERP).

1 EMPATHY

In Psycho’s iconic shower scene, the music assaults the listener. Marion stands in the shimmering white tub, placidly relaxing, as the water cascades down her body. The soothing white noise drenches the soundtrack, concealing all other sounds. With her back to the semitransparent shower curtain, Marion is unaware of a dark, undefined, looming silhouette. The shadow gets closer, taking the shape of an older woman. A hand with a menacing knife reaches up and rips the curtain. Marion screams in horror as the blade ruthlessly slashes her, again and again, as if tearing the very film apart—the comforting shower noise now turns to contorted musical madness, with violent violin gestures echoing Marion’s piercing cries for help. Through the music, we resonate with her frantic struggles—just as Marion, on-screen, cannot avoid her frightful destiny, we, the captive audience, cannot avoid sharing her agony, feeling her pain. Film music makes you feel what the characters feel. We feel scared when the character is scared, and brave when the character is brave. We feel angry, blue, joyous, lonely. We feel like we are inside the story, and part of the action. We feel what we hear. Listeners often ascribe such strong embodied responses to the music’s expressive power, but this brings about more questions than it answers: What musical forces make us empathize with the characters’ emotions? Are we genuinely experiencing these emotions or merely inferring their emotional states? What happens in our brains and bodies when we empathize with film characters? This first chapter begins to answer these questions by establishing a foundational framework for meaning construction: a model for empathy through film music. In fleshing out this model, we address relevant scholarship on the neural and bodily mechanisms underpinning the communication and interpretation of emotion through music, and present readings of various film scenes to probe the model and highlight film music’s extraordinary potential for meaning construction. Modeling Musical Empathy in Film

The model in Figure 1.1 outlines how empathy unfolds within the context of film music: (1) composers interweave musical gestures in the score, which function as correlates of vocal

DOI: 10.4324/9780429504457-2

Empathy

FIGURE 1.1

7

Process model of musical empathy.

and bodily gestures; (2) when listening to the music, audience members mirror these musical gestures via three sub-mechanisms (entrainment, subvocalization, and contagion); and (3) the physiological responses stemming from this internal simulation elicit in the audience the composer’s intended affective state.1 Within the context of this model, musical gestures are correlates of vocal and bodily gestures and postures, emerging from mapping vocal and bodily gestures and postures onto the musical space.2 Through this mapping, the associated expressive content of physical gestures or postures is not lost but translated.3 Whereas bodily gestures often translate to dynamic musical figures—e.g., melodic or harmonic movement—bodily postures translate to (relatively) static musical figures—e.g., single chords or notes.4 Because we share similar bodies and minds, our experience of living in our own bodies allows us to attune ourselves to the actions, sensations, intentions, and emotions of others. Therefore, central to this model is the presence of a music-based human mirror neuron system (MNS), which activates critical brain regions dedicated to perception and action while listening to music.5 Engaging the MNS in response to music, however, is not contingent upon musical knowledge or instrumental proficiency. While listening to music, musicians and nonmusicians alike sing along, tap the beat, mimic performative gestures (like air-guitar performances), and even interpret the emotional intentions behind the music.6 A musicbased MNS, therefore, allows for a powerful means of communicating emotions—music becomes a universal language. To capture music’s wide-ranging parametric variability, this model integrates three mirroring mechanisms: entrainment, subvocalization, and contagion. These mechanisms are analogous yet complementary, each responding to different musical parameters—subvocalization responds primarily to pitch, timbre, and contour; contagion to texture; entrainment to rhythm and tempo. Each also furnishes different empathic functions—while entrainment allows us to empathize with individual and group behavior, subvocalization works primarily at the individual level, and contagion at the group level. Lastly, this model rests on the counterintuitive premise that musically elicited affective states stem from our subjective experience of physiological changes in our bodies—simply put, bodily states cause emotions, rather than the other way around.7 As a result, by modulating our physiology—heartbeat, respiration, motor patterns—film music becomes a direct channel through which we empathize with film characters.

8

Empathy

Entrainment

By means of entrainment, we synchronize our bodies to temporally cyclic stimuli, to the regularities and periodicities of kinetic events in our environments. Research in neuropsychology provides ample evidence of premotor cortex activity during both production and perception of temporal stimuli, firmly situating entrainment as a subcomponent of the human MNS.8 And, as in all mechanisms related to the human MNS, we entrain to stimuli spontaneously and subconsciously, with motor responses manifested either overtly or covertly—for example, overtly when toes tap along, covertly when heartbeats match an external stimulus.9 Film composers draw on our capacity for entrainment to modulate our affective states.10 Because bodily changes trigger affective ones, and because cardiac and respiration muscle groups synchronize to a dominant sound, film soundtracks often foreground a character’s increased heart- or breathing-rate to modulate our physiology and subliminally heighten our alertness and stress response. Two examples, one from Midnight Express and one from Gravity, illustrate this common technique deployed within extramusical components of the soundtrack. In an early scene from Midnight Express, Billy, an American college student, is about to board the return flight from a vacation in Istanbul. He fears being caught smuggling two kilos of hashish strapped to his body. As he enters the security checkpoint, we hear his racing heartbeat. The guard studies Billy and sneers, “Bag!” Billy tenses and unhurriedly opens his shoulder bag, a bead of cold sweat rolling down his forehead. A second guard catches sight of the sweat and commands Billy to take off his sunglasses. Billy’s audible heartbeat increases further, as does ours. He knows his nerves are visible through his eyes, yet he daringly takes off his sunglasses and stares at the guards, trying not to look away. Throughout this nerve-wracking scene, the thumping of his heart comes into the foreground, pushing other soundtrack elements to the background, prompting us to entrain to Billy’s heartbeat and experience his apprehension firsthand. In Gravity, the soundtrack often foregrounds the protagonist’s breathing, prompting us to entrain to it, thereby modulating our affective states. Outside the Explorer Space Shuttle, Dr. Ryan Stone carries out her first repair mission. Her bulky white spacesuit is strapped to a robotic arm, preventing her from drifting away into space. She safely floats close to the shuttle, her measured breathing allowing her to focus on her work. Something flickers in her eyes—a cloud of debris hurtling toward the shuttle, far too fast. A bulky piece of debris hits the right wing, causing the robotic arm to flail uncontrollably. Another piece hits the robotic arm, causing it to spin away from the Explorer at an incredible speed, Ryan still strapped to it. Stars and objects orbit wildly in her field of vision. She must detach from the robotic arm

FIGURE 1.2

Midnight Express. Billy’s heartbeat. [00:04:30]

Empathy

FIGURE 1.3

9

Gravity. Audible increase in Ryan’s cardiorespiratory rate. [00:13:30]

before it carries her too far from the shuttle. Ryan panics—we hear her breathing quicken. Trembling, she unhooks the clasp, finally thrusting herself away from the robotic arm. Her eyes sift through the black and incalculable void, glimpsing the shuttle, now shriveled into a tiny dot in the distance. Her breathing speeds up further, so does ours. As the darkness swallows her, she calls out for help: Explorer, do you . . . do you copy? Houston, do you copy? Houston, this is Mission Specialist Ryan Stone. I am off structure, and I am drifting. Do you copy? Anyone . . . ? Anybody . . . ? Do you copy? Please copy. Please. At the climax of this angst-ridden scene, the soundtrack ties her speech’s syntactic disintegration to the increase in her breathing rate, compelling us to embody Ryan’s terrifying experience by physically affecting us through entrainment.11 Entrainment is not limited to synchronizing to others’ bodily movements or audible vital signs—cyclic patterns in the music also engage listeners through entrainment.12 Entraining to the music’s tempo triggers physiological changes that, in turn, elicit affective ones.13 Film composers skillfully draw on this phenomenon, prompting us to empathize with the characters, especially while they experience moments of high arousal. The music in Lord of the Rings: Return of The King grips us and makes us feel what the characters feel. The opening scene flashes back to Sméagol’s birthday, fishing with his cousin Déagol at the idyllic River Anduin. Déagol hooks a mighty fish, which hauls him overboard. As he lets go of the line, a golden ring lying in the silt catches his eye. Déagol climbs out of the water, holding the glittering ring in his palm. The ring now reflects in Sméagol’s eyes, who entreats, “Give us that, Déagol, my love! Because it is my birthday, and I wants it.”

FIGURE 1.4

Lord of the Rings: Return of The King. Déagol’s thumping heartbeat. [00:02:30]

10

Empathy

FIGURE 1.5

American Animals. Spencer thumping a heartbeat rhythm. [00:57:30]

Déagol ignores him, a smirk on his face. Driven by the ring’s destructive forces, they turn on each other, fighting for possession of the ring. Sméagol tries to snatch the ring from Déagol’s hand, again and again. Sensing that Déagol will not yield, Sméagol goes for his neck. In the soundtrack, the ring’s hum blends with a musical rendition of an elevated heartbeat. As Sméagol chokes Déagol to death, the thumping in the music decelerates and ultimately ceases. Sméagol hisses, “My precious”, takes the ring as his own, and disappears. The underscoring in this spellbinding scene prompts us to subliminally entrain to Déagol’s heartbeat, allowing us to empathize with his distressing final moments and experience the power of the ring to turn beloved cousins against each other in cold-blooded conflict. In a scene from American Animals, the initial heartbeat-mimicking sounds become a stylized rendition of a heartbeat in the underscoring. Four college students plot to steal valuable books from their university’s Special Collections Library. On the day of the heist, Spencer and Warren, unconvincingly disguised as older men, wait in a living room for Chad to arrive. Spencer sits at a table, expressionless, tapping a heartbeat using two porcelain animals. Warren paces back and forth, constantly checking his wristwatch. Spencer’s unremitting tapping puts Warren on edge, causing him to shout at Spencer to make him stop. Although the diegetic thumping ceases, a heartbeat rhythm emerges as non-diegetic music. Chad arrives, also disguised as an older man. The three drive at a snail’s pace through a residential area, yet the musical rendition of an unrelenting heartbeat persists in the soundtrack. Eric, the fourth man, stands waiting at the side of the road. Their van pulls up, and he gets in. The four men sit in tense silence, while the musical heartbeat resonates in the soundtrack. They arrive at the university campus and step out of the van. As they stroll through the open quad toward an imposing mock-colonial building, the Special Collections Library, the stylized musical rendition of the heartbeat becomes buried in other musical layers. The initial transference of the heartbeat rhythm in the sound design—from diegetic (Foley) sound to non-diegetic music—gives us an extended window of time to subliminally entrain to it, thereby ensuring we will empathize with the characters’ apprehension and experience their elevated tension in anticipation of the heist. Our tendency to entrain to musical stimuli’s temporal regularities also allows directors to establish a consistent pace in scenes with irregularly paced visuals. The ‘sensual haircut scene’ from Edward Scissorhands exemplifies this phenomenon. All the housewives line up with their dogs in a neighbor’s backyard. Joyce gapes as she holds Kisses on top of a table and Edward grooms the toy poodle into a tiny, unique work of art. Fascinated, she flirts, “Is there anything you can’t do, Eddy? I swear, you take my very breath away. Have you ever cut a woman’s hair? Will you cut mine?” As she offers herself on a chair in front of Edward,

Empathy

FIGURE 1.6

11

Edward Scissorhands. Joyce indulges under Edward’s blades. [00:49:30]

a sensual habanera sets in the music. He confidently swivels her head, studies her hair. She flushes in anticipation. With calculated precision, he unleashes his skills—snipping here, layering there. She luxuriates under his unwavering blades, eyes closed, a sensual smile on her lips. The exhilarating music follows Edward’s swift and rousing movements, brushing by the visually slow moments depicting Joyce’s ecstatic response, moments that threaten to undermine the scene’s whirling relentlessness. He continues, cutting dangerously close. She twirls her feet in pleasure. As he makes a final cut and steps back, satisfied, the music eases back into the habanera. Glowing, Joyce purrs, “That was the single most thrilling experience of my life.” Via entrainment, the relentless pace of the music overrides the pace of the visuals, allowing us to experience the rousing intensity of the moment, bringing us along with Joyce to a riveting climax; subsequently, the habanera takes us to a pleasurable repose, helping us recognize how Edward’s sharp and strong hands helped transform Joyce’s persona as much as her appearance, revealing the latent seductress within. Subvocalization

Laughter, screams, cries, shouts, gasps, sighs—all are part of an extensive repertoire of vocal gestures expressing innumerable affective states, including happiness, anger, sorrow, triumph, awe, and frustration. These vocal gestures subliminally activate the listener’s vocal musculature and neural substrates associated with producing those gestures via a mechanism known as subvocalization.14 However, subvocalization is not limited to mirroring isolated vocal gestures. Recent research shows that singing—which straddles the boundaries between verbal and nonverbal gestures—also triggers subvocalization.15 As a result, because nonverbal vocalizations and singing share the same production mechanisms and expressive purposes, these modes of communication also share the associated physiological and affective states resulting from subvocalization—when we mimic vocal sounds, we feel firsthand the emotion these sounds project. When subvocalizing, we channel a signal’s acoustic signatures—subtle yet salient nuances in a signal’s sonic parameters—through our vocal apparatus, (re)enacting the associated physiological processes that trigger them and thus educing the associated emotions. For instance, vocalizations expressive of anger feature anger’s characteristic acoustic signatures—harsh timbre, angular contour, loud dynamic level, dissonant overtones. Similarly, vocalizations expressive of tenderness feature that emotion’s contrasting characteristic acoustic signatures—mellow timbre, smooth contour, moderate dynamic level, consonant overtones.16 Acoustic signatures are therefore foundational to the mechanism of musical empathy, transforming meaningless sonic signals into gestures with expressive power. Film

12

Empathy

FIGURE 1.7

Gladiator. Maximus mourns the death of his family. [00:42:50]

music exploits the expressive power of acoustic signatures embedded in vocal music to great effect, especially to construct an empathic bond between film characters and audience members. The human voice plays a pivotal role in Gladiator’s musical score, prompting us to empathize with the protagonist’s affective states by engaging our MNS via subvocalization. Maximus returns home from combat. At the sight of thick black smoke rising over the horizon, he gallops homeward in a frenzy. He arrives at his worst nightmare—vineyards destroyed, the earth scorched, houses smoldering—and collapses off the horse. The music emerges with a rendition of a solo vocal lament over a low drone. Pulling himself up, he staggers past the rubble, anticipating his greatest fear. As he catches sight of two crucified and charred bodies—his wife, his son—he howls, tormented, sinking into despair. The music penetratingly depicts Maximus’s mourning throughout the scene using musical representations of vocal and nonverbal sounds. The solo voice texture reflects the intimate grief of an inner psychological space—the timbre of closed-mouth vocalizations conveys a suppressed emotion, and the melodic contour suggests the visceral representation of an animal lament.17 While listening to the music, we subliminally engage our MNS through subvocalization, which compels us to resonate with Maximus’s suffering. Subvocalization extends even further. Research in auditory neuroscience broadens the notion of subvocalization to include responses to a wide range of auditory stimuli, including instrumental music.18 When we hear instrumental music, we mimic it internally, emulating its sonic qualities within our vocal capabilities—that is, we resonate strongly with vocal and instrumental gestures, provided we can reproduce them vocally. Instrumental music’s expressive capacity is thus scaffolded by acoustic signatures stemming from vocal gestures.19 However, since our auditory perception attunes mainly to human vocalizations, vocal music elicits more powerful reactions than instrumental music; therefore, instrumental gestures approximating voice-like characteristics will trigger a more pronounced subvocal response.20 The music in 15 Minutes maps the characters’ bodily and vocal gestures onto instrumental ones. The film chronicles the chaotic journey of Emil and Oleg, two Russian ex-convicts traveling to the U.S. to collect their share of a heist. The musical underscoring reacts in tandem with their experiences, subliminally guiding our emotions, constructing an affective bond that allows us to empathize, yet not sympathize, with these utterly unlikable characters. Two scenes exemplify this phenomenon by bringing back a motivic idea used at the beginning of the film during the main titles sequence for character construction, albeit radically transformed to communicate two widely different affective states.

Empathy

13

Early in the film, Emil and Oleg pay an unexpected visit to their old partner, Milos, who has settled for a modest yet peaceful life in New York with his wife. Oleg, obsessed with becoming a Hollywood filmmaker, captures these moments with a handheld camera. As Emil demands his rightful share of the heist, he learns that Milos has spent all the money. Emil furiously erupts, stabbing Milos and strangling his wife. Confused by his own murderous impulse, Emil zigzags through the modest apartment, sobs, hunkers down, gesticulates as if trying to explain what has just happened. In the background, wild musical gestures performed on a violin echo these unsettling events—brusque attacks, wide and unpredictable leaps, noisy and coarse timbres. Emil ransacks the old kitchen cabinets, tossing empty cans and bottles, looking for something. Finding a metal container with acetone, he turns to the bodies. Oleg, who has been filming the havoc, grunts, “Aw . . . Bohemian barbecue.” Emil closes the curtains, shrouding the room in darkness. Transformed beyond recognition, the playful motivic idea heard at the beginning of the film now allows us to experience Emil’s bewilderment while accentuating his wrath.21 The underscoring’s musical gestures, featuring acoustic signatures characteristic of anger, prompt us to empathize with the character’s affective state—Emil’s fury—via subvocalization while being horrified by his actions. Toward the end of the film, Oleg is fatally shot. Sprawled on the ground, video camera in hand, he seizes the opportunity to capture a realistic performance of his own death. He slowly pans from the Statue of Liberty toward a close-up shot of himself dying. A violin enters the musical texture, but this time gently following Oleg’s quiet closing credits, anticipating the imminent mourning. Faithfully shadowing Oleg’s death, the motivic idea now features the acoustic signatures of sorrow—smooth contour, narrow range, regular rhythm, predictable harmonic flow. Once again, the underscoring’s musical gestures prompt us to empathize with the character’s affective state—now Oleg’s suffering—via subvocalization.

FIGURE 1.8

15 Minutes. Emil’s murderous wrath. [00:14:30]

FIGURE 1.9

15 Minutes. Oleg films his death. [01:50:00]

14

Empathy

Among the many acoustic signatures, timbre is arguably the most efficient in triggering an affective response.22 Our perception of timbre is shaped by a tone’s harmonics, their interaction affecting the composite sound’s degree of sensory consonance and dissonance.23 For instance, the sounds of a flute, ocarina, or French horn lack inharmonic partials (those that deviate from whole multiples of the fundamental frequency), resulting in the instruments’ characteristic ‘mellow’, ‘warm’, ‘smooth’ timbre. In contrast, the sounds of a trumpet, oboe, or most percussion instruments feature copious inharmonic partials, resulting in their characteristic ‘harsh’, ‘rough’, ‘noisy’ timbre.24 Consequently, the presence and intensity of inharmonic partials influence the associations we form between timbres and affective states—dull timbres with sadness, brash timbres with anger.25 Film composers are keenly aware of timbre’s effectiveness in triggering a rapid emotional response in the listener, and hence are keen on choosing specific timbre(s) when orchestrating musical gestures. In the two examples from 15 Minutes, the widely different violin timbres, resulting from different performance techniques, convey radically different emotions. However, while some instruments, such as the violin, can produce a broad range of timbres, others are more consistent in their timbric variability, making them film music staples of certain affective states. In Jurassic World: Fallen Kingdom, the music projects honor and loyalty onto a dinosaur. It introduces a French horn, whose natural timbre meets the necessary acoustic signatures—high relative amplitude of the fundamental, low-frequency formants, low perturbation, few inharmonic partials.26 Toward the film’s end, Owen attempts to persuade Blue, the last surviving female velociraptor, to be caged in a ‘sanctuary’. Outside the Lockwood Mansion, Blue’s birthplace, Owen treads down the steps toward her. Having trained Blue from birth, he recognizes her noble nature and assures others, “It’s okay. She won’t hurt us.” The underscoring to Owen’s words features a calming melody on a French horn performing in its most comfortable register. Owen extends his hand to caress Blue and mutters, “Hey, girl, come with me. We’ll take you to a safe place, okay?” Blue chitters as she follows Owen’s gaze to a large, barred enclosure. Breathing heavily, she backs away from his hand and locks eyes with him as though saying goodbye, and, rejecting a life in captivity, she sprints into the forest to join the other dinosaurs. The music contributes to anthropomorphizing Blue, prompting us to empathize with her while reminding us that she is a loyal and honorable being. Moreover, by bestowing Blue with human traits (such as nobility), the music foreshadows the film’s final message: humans must learn to live with their environments and must embrace the “humanity” of other animals.

FIGURE 1.10

Jurassic World: Fallen Kingdom. Owen recognizes Blue’s noble traits. [01:55:00]

Empathy

FIGURE 1.11

15

Psycho. Marion’s piercing screams. [00:47:30]

Subvocalizing instrumental timbres that emulate a scream or shriek would be borderline painful—screaming requires us to tense our vocal folds, resulting in a sound that embeds many inharmonic partials. This brings us back to Psycho’s shower scene. As the shattering silver blade flashes and slashes Marion, again and again, the unremitting violin acciaccaturas provoke forceful frictions between the bow and the strings, like when sharpening a knife, generating inharmonic partials, thereby infusing this instrumental gesture with a scream’s characteristic acoustic signatures. When subvocalizing these high-pitched strident musical gestures, we empathize with the character’s frantic struggle, we feel her pain. What is true for complex tones is arguably true for harmonic constructs, particularly regarding the different degrees of embedded consonance and dissonance.27 This parallel, between the sensory consonance and dissonance of both complex tones and harmonic constructs, begins to reveal the source of our deeply rooted associations between harmonic consonances and dissonances and our affective states.28 Film composers may thus infuse a score with harmonic consonances and dissonances to reflect narrative developments, using music as an off-screen narrator. Another scene from Jurassic World: Fallen Kingdom nicely exemplifies the compositional decision to use consonant and dissonant musical gestures aligning with and emphasizing the affective states suggested by the narrative. A young character, struggling to escape a dinosaur, grabs a rescue ladder dropped by a helicopter and begins climbing. The dinosaur also catches on to the ladder, potentially bringing down both the young man and the helicopter—the underscoring to this tense moment is saturated with dissonant harmonic constructs. As the dinosaur fortuitously releases the ladder, the character enthusiastically celebrates narrowly escaping the threat—the underscoring to this joyful moment turns to consonant harmonic constructs, particularly major triads. Only seconds later, however, highly dissonant sonorities underscore a shocking turn of events: an enormous leviathan erupts from the water and swallows the character whole.29

FIGURE 1.12

Jurassic World: Fallen Kingdom. Near-successful escape. [00:05:45]

16

Empathy

Contagion

Musical contagion is rooted in ecological and evolutionary functions of prosocial interaction and cooperation, such as ensuring task efficiency or the welfare of ingroup members.30 Synchronizing with others through music helps develop ingroup bonding and allows us to collectively attain signal amplification by combining our individual sounds into a louder and stronger whole.31 For example, singing or performing in groups engages both producers and listeners, leading to the social cohesiveness, bonding, and affiliation typical of sporting events, religious ceremonies, and political rallies. Within these environments, and like behavioral contagion, musical contagion is automatic and subliminal, stemming from our evolutionary tendencies toward developing social networks and ingroup cohesion.32 Film composers harness these powerful evolutionary and ecological tendencies. By introducing voluminous choirs or sizable instrumental ensembles where all members contribute to a unified and cohesive whole, composers prompt us to engage in musical contagion, subliminally leading us to align with the collective and become co-participants of the film’s events. As illustrated in the earlier example from Gladiator, the solo voice engages our MNS through subvocalization, allowing us to empathize with the protagonist’s affective states. In other moments in the film, the choir in the underscoring imbues Maximus with the power he will exert as a leader of gladiators, engaging our MNS through contagion. The film thus weaves through the various dualities of the protagonist’s personae—the intimate versus the communal, the emotional versus the impassive, the desire for peace versus the inevitability of war—and maps this duality in the music. At a vital moment in the film, Maximus leads three untrained gladiators to a decisive victory in a brutal fight. As they stride into the magnificent Colosseum, a massive wall of sound overtakes them—rowdy roars of ferocious lions, vicious grunts of their opponents, swathing ovations from the agitated arena. The battle is on, unleashing a chaotic flurry of dust and blood. Maximus takes control, leading his gladiators, joining forces, battling together. The musical underscoring in these bonding moments features the confident sound of a male choir in unison with the orchestra, with an arrangement foregrounding the muscular sound of double basses and brass instruments. The bustling energy at the arena mounts as Maximus wades through his opponents, slashing through them heroically. A staggering spectacle. Maximus glances around, all his opponents defeated. The arena is a graveyard of body and animal parts. As the triumphant gladiators exit the arena, a crowd of thousands stomps and cheers, “Maximus! Maximus!” By means of contagion, the music places us alongside the crowd, partaking in the gladiators’ victory—the

FIGURE 1.13

Gladiator. Maximus leads untrained gladiators to victory. [01:28:00]

Empathy

FIGURE 1.14

17

Africa: The Serengeti. The newborn calf joins the herd. [00:35:10]

great number of voices constructs a colossal virtual performing space, and the homorhythmic movement of these voices projects (and instills in us) a sense of vigorous coalition. Using a choir to propel a sense of collective identity is not limited to stories about people. In animal documentaries, for example, filmmakers deploy various cinematic techniques and tropes that contribute to constructing an anthropomorphic narrative. In a sequence from the documentary Africa: The Serengeti, we learn from an off-screen narrator that “To survive in the midst of predators, calves must be able to stand within minutes of birth.” The sequence begins with a struggling newborn wildebeest, surrounded by an apprehensive herd, prompting the calf to stand—in the music, a distressed solo female vocalist in the high register sounds in counterpoint to a ceremonial humming of a male choir in the low register. While we further learn that “Those unable to stand must be abandoned”, the calf is shown near ferocious predators waiting for the herd to leave. In a last impulse of life, the calf finally manages to stand. The herd jauntily ‘cheers’ and once more surrounds the calf, depicting a triumphal, exultant moment of ingroup cohesion. The music maps this event by anthropomorphizing the herd’s sentiment of jubilation—female and male voices now join forces, aligning their singing homorhythmically, joyfully spanning a wider register, swelling with active melodic contours. Along with the cinematography, the music in the scene engages our MNS—first via subvocalization and then contagion—prompting an empathic response in the listeners, a response grounded in their own human phenomenological experience yet funneled into a sympathetic communal affiliation with the herd. Although introducing music featuring a choir seems the most effective means to elicit contagion, many excellent examples feature large orchestras instead. In Star Wars: The Rise of Skywalker, Poe and the Resistance battle the First Order over Exegol. Outnumbered and overpowered, Poe hopelessly yields in defeat, calling out, “My friends. I’m sorry. I thought we had a shot. But there’s just too many of them.” The music underscoring Poe’s words features a single French horn, prompting us to empathize with his (and the Resistance’s) noble and honorable goals. The voice of Lando Calrissian suddenly breaks in, signaling a massive fleet of ships arriving from across the galaxy, spurring on the Resistance: “But there are more of us, Poe. There are more of us!” With an outburst of orchestral forces underscoring Lando’s words, the music also says, “There are more of us!”, and via contagion, it brings us, the audience, into the film’s action as participants in the battle alongside the Resistance, subliminally assuring us that ‘we’ now have the power in numbers to defeat the First Order.

18

Empathy

FIGURE 1.15

Star Wars: The Rise of Skywalker. Massive fleet of ships arrives. [01:53:30]

Coda

Film music puts us right there with the characters on the screen. It makes us feel what they feel. By drawing on our innate capacity to empathize with others, it places us in a shared, intersubjective, intercorporeal space that allows us to attune ourselves to the characters’ emotions, sensations, and intentions.33 In this chapter, I introduced a model for musical empathy in film. This model, like its social empathy counterpart (outlined in Appendix I), unfolds as a mosaic of subcomponents mediated by the mirror neuron system: (1) film composers map the essential features of vocal and bodily gestures onto the musical space, (2) we subliminally mirror the original vocal and bodily gestures via subvocalization, entrainment, and contagion, and (3) this mirroring mechanism elicits in us the intended physiological and affective responses. Ultimately, we redirect this empathic response toward constructing an interpretation of the film’s narrative.34 Understanding musical empathy in terms of its constituent mechanisms—subvocalization, entrainment, contagion—allows us to recognize that the music in Psycho’s iconic shower scene was carefully designed to assault the listener. Whereas the stylish visuals, with extreme close-ups and swift edits, only imply violence, the shrieking violin gestures ratchet up the tension and subliminally force us to mirror the character’s pain. While we do not see the knife stabbing Marion, we feel the psychotic violence through the piercing music. Notes 1. This musical empathy model substitutes several components of the social empathy model I propose in Appendix I in this volume. As with the proposed model for social empathy, the physical and psychological attunement this musical empathy model purports has not been empirically verified as a whole; nevertheless, much recent research in auditory neuroscience and allied fields supports the individual mechanisms contributing to the music- or sound-borne intersubjective mirroring of physiological and affective states. 2. Philosophers and musicologists have long recognized such seamless cross-domain mapping of musical gestures and their associated affective states: Barthes (1985) alludes to the expressive interplay between the music and the body, and characterizes musical gestures as “figures of the body, whose texture forms musical signifying” (p. 306); Hatten (2004) points us to the intermodal (or rather amodal) nature of musical gestures and their communicative function, defining them as “emergent gestalts that convey affective motion, emotion, and agency” and noting that “the basic shape of a gesture is isomorphic and intermodal across all systems of production and interpretation” (p. 109); and similarly, using language that traces musical gestures as emerging from neural and physiological states, Lidov (2005) notes that “the notion of expressive gesture interpreted as the surface form of an underlying neurological function gives us a logical connection between

Empathy

3.

4. 5.

6. 7.

8.

9.

10.

11.

19

somatic and musical experience, one which shows a basis for cause and resemblance in the relation of music to feeling” (p. 152). While vocal and bodily gestures and postures may have a social or cultural origin, this chapter only considers musical gestures and postures stemming from biological or physiological states. Later chapters consider gestures that emerge from social milieus and those grounded in symbolically motivated conventionalized movements, which result in arbitrarily established and conventionalized figures (accompaniment patterns, metrical parameters, rhythms, etc.) and which rely on the listener’s cultural and gestural competency to be understood. Godøy (2003) examines the connection between gestural imagery and musical imagery, arguing that gestural images are essential in triggering and sustaining mental images of musical sound. Much research draws on experimental evidence to extrapolate a specialized music-based mechanism from a more generalized MNS. For instance, Koelsch et al. (2006) and Menon and Levitin (2005) show we engage critical regions of the MNS (the premotor cortex and insula) during music listening. Extending similar research to a potential mechanism of emotional communication through music, Overy and Molnar-Szakacs (2009) explain that, while listening to music, “interactions between the [MNS] and the limbic system may allow the human brain to ‘understand’ complex patterns of musical signals and provide a neural substrate for the subsequent emotional response” (p. 490). For further insights on a music-based MNS see Gridley and Hoff (2006); Molnar-Szakacs and Overy (2006). Overy and Molnar-Szakacs (2009) show that nonmusicians “activate the MNS during music listening when mapping [musical stimuli] onto basic, nonexpert musical behaviors that they are able to perform such as singing, clapping and tapping” (p. 494). Much research supports this proposition. For instance, Molnar-Szakacs and Overy (2006) cites research showing that music induces facial expressions in listeners which in turn elicit affective states, leading them to conclude that “the perception of emotion in music may arise in part from its relation to physical posture and gesture” (p. 238). Large and Snyder (2009), for example, argue that the “perception of pulse and meter result from rhythmic bursts of high-frequency neural activity . . . [which] enable communication between neural areas, such as auditory and motor cortices” (p. 46). Given the necessary temporal accuracy necessary when investigating entrainment, most studies draw on MEG or EEG. For further insights on the neuropsychology of entrainment see Doelling and Poeppel (2015); Fries (2005); Grahn and Brett (2007); Janata and Grafton (2003); Nozaradan et al. (2011); Will and Berg (2007). Seeking to understand the covert response to entrainment, Fujioka et al. (2012) used magnetoencephalography (MEG) to measure the participants’ neural oscillations while they passively listened to rhythms; their results show that, although participants remained motionless, their neural oscillations (Beta waves) synchronized to the rhythmic stimuli. In addition, neuroimaging research shows that entrainment to rhythmic periodicities (including biological rhythms) recruits the perceptual, motor, and sensorimotor cortical areas (Grahn & Rowe, 2009; Zatorre et al., 2007). For further insights on entraining to heartbeats see Anishchenko et al. (2000); Lorenzi-Filho et al. (1999). To date, however, there is no conclusive evidence regarding the relationship between the music’s tempo and the listeners’ heartbeat rate. A thrilling scene in Barbarian offers a compelling example of how the sound design can potentially elicit increased heart and breathing rates—as Tess descends a dimly lit staircase into the underground tunnels of a rental house, her heart and breathing rates audibly intensify. [00:39:00] Although there is no conclusive evidence that individuals spontaneously entrain precisely to the music’s tempo—that is, below the level of conscious attention—most studies suggest that the music’s tempo subliminally modulates our behavior. For instance, Leman et al. (2013) suggest that “some music is activating in the sense that it increases the [walking] speed, and some music is relaxing in the sense that it decreases the [walking] speed” (p. 1). Much like listening to (and mirroring) musical gestures by others causes in us physiological changes that induce affective states, producing our own musical gestures also causes physiological changes that modulate our own affective states. We can (and often do) utilize this mechanism for self-regulation. The beginning moments in the scene from Gravity illustrate the use of entrainment and subvocalization as a means for self-regulation. Debris dangerously crashes at high speeds, shattering parts of the station, but Ryan has no option but to continue untangling the parachute ropes. In an attempt to calm herself, she begins humming a peaceful melody. Her humming serves her well as an affective self-regulatory mechanism, keeping her thriving in a dangerous situation.

20 Empathy

12.

13.

14.

15.

16.

17. 18.

19.

20.

However, her voice and the music vie for control of the narrative, and ultimately, the relentless underscoring, with its ever-increasing intensity and agitation, overpowers the audience’s senses and forces us not to be soothed by her tranquil lullaby. Sato et al. (2012), for example, provides evidence that listeners “unconsciously changed their respiration timing to coincide with the music track” (p. 255). Analogously, Juslin et al. (2010) argue that the “rhythm of the music interacts with an internal body rhythm of the listener such as heart rate, such that the latter rhythm adjusts towards and eventually ‘locks in’ to a common periodicity” (p. 621). For example, Husain et al. (2002) show that exposure to fast tempi increases arousal and tension. Similarly, in a related study on brain stem reflex—which controls processes of the autonomous nervous system such as pulse, respiration, heart rate, and skin conductance—Juslin and Västfjäll (2008) explored the processes whereby music induces emotions. They identified tempo as the most significant parameter in a modulating effect, and as the parameter triggering the broadest range of emotional responses. For an in-depth exploration of the psychophysiological responses to musical tempo and entrainment see Etzel et al. (2006); Khalfa et al. (2008); Levitin (2006); Scherer and Coutinho (2013); Trost and Vuilleumier (2013); Van der Zwaag et al. (2011). Numerous studies record covert subvocalizations at the musculoskeletal level using laryngeal electromyography (e.g., Brodsky et al., 2008) and at the neural level using functional magnetic resonance imaging and single-cell recordings (e.g., Harris & De Jong, 2014; Mukamel et al., 2010). Additionally, researchers use transcranial magnetic stimulation to disrupt brain activity in motor areas associated with subvocalization; Lima et al. (2016), for example, note that impairing subvocalization also impairs speech discrimination. For further insights on subvocalization see Bestelmeyer et al. (2014); McGettigan et al. (2015); Warren et al. (2006). Subvocal responses to human singing have been widely observed at both the musculoskeletal and the neural levels. Lévêque et al. (2013) note that “the perception of a human-produced sound like the singing voice [induces] motor resonance via interactions between the auditory and vocal systems” (p. 1), and Callan et al. (2006) found that “neural processes underlying both perception and covert production of singing and speech activate overlapping brain regions” (p. 1334). Di Stefano et al. (2022) examine the perception and aesthetics of musical consonance and dissonance and reviews relevant scholarship on the topic. They present three hypotheses (vocal similarity, psychocultural, and sensorimotor) while exploring the biological underpinnings of the attraction-aversion mechanisms triggered by consonant and dissonant stimuli. Sachs (1962) speculates that particular melodic contours derive from animal howls or wails; he identifies examples in Western classical, Russian, Australian aboriginal, and Lakota (Sioux) music. Much research explores subvocalization triggered by nonvocal and nonmusical sounds. In a systematic review, Lima et al. (2016) argue that “because of the multifaceted and flexible nature of vocal production [responders] generate sensorimotor estimations of different properties of sound” (pp. 538–539). Using fMRI, Koelsch et al. (2006) observe that instrumental music activates the brain region responsible for both perceiving and executing vocalizations; they note this activation in the premotor cortex area responsible for movements in the larynx, even when participants did not exhibit overt movements. From a speculative, albeit empirically informed perspective, Cox (2001) conflates vocal and instrumental subvocalization, remarking that “when others speak or sing, we understand these sounds partly in terms of our own experience of making the same or similar sounds. When others make sounds on musical instruments . . . we understand these human-made sounds . . . in terms of our own vocal experience, by way of subvocalization” (p. 201). Other scholars acknowledge a mirroring mechanism without directly addressing a mirror neuron system; Heidemann (2016), for instance, notes that “In listening to vocal music, we may involuntarily mirror the actions we imagine the performer undertaking” (par. 1.1). Upon surveying existing scholarship, Lima et al. (2016) conclude that passive music listening also activates neural substrates responsible for subvocalization and that, more generally, the “supplementary and pre-supplementary motor areas play a role in facilitating spontaneous motor responses to sound” (p. 527). This phenomenon, which is consistent with findings related to the MNS, has been explored in relation to vocal versus instrumental melodies (Watts & Hall, 2008) and in relation to the speed of auditory processing of vocal versus nonvocal sounds (Agus et al., 2010). Lévêque and Schon (2015), for example, found increased activity in the auditory cortex to vocal melodies; they note

Empathy

21. 22.

23. 24. 25. 26. 27.

28.

29.

30.

31.

21

that “hearing a (human) singing-voice involves more strongly the sensorimotor system than hearing the same melody played with a non-human timbre”, concluding that this increased activation is due to “a facilitated matching between the perceived sound and the participants motor representations” (p. 58). For further insights on the subvocalization of instrumental versus vocal timbres see Wilson et al. (2004). This example supports Hurley’s (2008) assertion that “hearing anger expressed increases the activation of muscles used to express anger” (p. 10). Timbre requires the least amount of time to unfold, while other acoustic signatures, such as contour, harmonic progression, or rhythm, necessitate much longer timeframes. Gjerdingen and Perrott (2008), for example, identified that participants could classify a piece’s genre (which is primarily defined by its timbre) within a 250- to 475-millisecond timeframe. While sensory dissonance—the physiological phenomenon of ‘beating’ resulting from exposure to two pitches—can be accurately quantified as manifested in the cochlea, musical dissonance is an evaluative notion contingent upon compositional style and musical vocabulary. For the acoustic signatures of various instruments see Juslin and Laukka (2003). For research on the relationship between timbre and emotion in music see Balkwill and Thompson (1999); Gabrielsson and Juslin (1996). Originally, the French horn did not feature valves and thus was only capable of producing the notes of the harmonic series. In an experiment measuring degrees of skin conductance in response to consonant and dissonant harmonic constructs, Winold (1963) found that dissonance triggers more significant sweating than consonance. Studies attempting to trace the ontogenetic origin of our response suggest that our aversion to dissonances and our preference for consonance is innate. When presented with a consonant and a dissonant stimulus, for example, infants attend to consonant stimuli for longer times, suggesting that they are preferentially attuned to consonances rather than to dissonances (Crowder et al., 1991; Masataka, 2006; Trainor et al., 2002; Zentner & Kagan, 1998). However, Plantinga and Trehub (2014) question this research and note that their findings are “inconsistent with innate preferences for consonant stimuli” (p. 40), suggesting that our preference related to consonance/ dissonance must be a learned response. Numerous hypotheses contribute to this line of thought, some grounded in perception—e.g., the presence of beating (or roughness) when hearing two dissonant pitches triggers spectral masking, hindering our capacity to perceive the individual stimuli and in turn causing a feeling of irritation (Huron, 2006)—some grounded in ecological or evolutionary theories—e.g., sensory dissonance is characteristic of warning calls and intimidation vocalizations in many species, and hence associated with potential danger (Ploog, 1992)—and some grounded in neuroscientific observations of functional covariations of brain regions—e.g., sensory dissonance activates the parahippocampal gyrus, an area implicated in the processing of stimuli with negative emotional valence (Blood et al., 1999; Koelsch et al., 2006). In the example from Psycho, the initial shrieking gesture expands toward the lower register, introducing major sevenths and minor seconds toward forming harsh harmonic dissonances. The music thus unfolds from dissonance within a complex tone toward dissonance within harmonic constructs. Lima et al. (2016), for example, note that “some sounds, owing to their rhythmical patterns [i.e., entrainment] or their social and motivational salience [i.e., contagion], elicit motor responses, such as singing, tapping, dancing, or vocal alignment . . . Such propensity to respond to rhythmic, social, and emotional auditory information might promote social convergence, learning, coordination, and affiliation” (p. 537). Within the context of the BRECVEMA model, Juslin (2013) describes ‘emotional contagion’ in music as “a process whereby an emotion is induced by a piece of music because the listener perceives the emotional expression of the music, and then ‘mimics’ this expression internally” (p. 241). While the main thrust of BRECVEMA draws on an evolutionary perspective, the model presented in this chapter draws on embodiment and social cognition. Overy and Molnar-Szakacs (2009) note that “whether making entirely different musical contributions to weave a musical texture, or all producing exactly the same sounds, the whole is much greater than the individual parts, from a choir to a drum circle to the stadium bleachers. The emerging sound is a group sound, almost ‘larger than life’, created by a sense of shared purpose”

22 Empathy

(p. 495). They also warn us, “when group music-making reaches a certain level of cooperation and coordination, the sense of shared purpose and togetherness can be extraordinarily powerful and even threatening” (p. 495). See also Merker et al. (2009); Phillips-Silver et al. (2010). 32. In a systematic review of the literature on embodied simulation, Juslin and Västfjäll (2008) identified ecological and evolutionary functions of a music-based MNS, including enhancing group cohesion and social interaction, learning, and emotional contagion. 33. Hoeckner et al. (2011) introduced the idea of empathy to film musicology, yet they do not propose a model for musical empathy or distill the various mechanisms involved. 34. Musical empathy represents a compelling resource for composers and proves foundational to theorizing more broadly about meaning construction in film music; however, as the following chapters will discern, it is but one of many sources we draw upon when constructing interpretations.

2 CONTAINER Schema

In The Truman Show, the music (de)constructs a fabricated reality, the ‘show’ that is Truman’s life. It is nighttime. Truman is fast asleep. A gentle and melancholic piano piece underscores a close-up shot of his peaceful face. Christof, the show’s creator, approaches a giant ON-AIR monitor broadcasting Truman to the world and reaches out with his arm, virtually caressing Truman’s face. Truman twitches, almost as if feeling Christof’s presence. As the camera pans away from this intimate moment to show the entire production studio, a pianist comes into sight—to our surprise, he is playing the very music we hear. By disrupting the storyworld’s sonic boundaries and transferring sounds within them, the soundtrack reveals Christof’s power and control over the fictional narrative Truman inhabits and over our interpretation of that narrative. Films immerse us in their storyworlds. They surround us with music, sound effects, and dialogue—all distinct components of their fictional reality. Every so often, these components become rearranged, freed and unbound from normative paradigms and sound design conventions, prompting us to construct metaphorical interpretations of the film’s narrative. In this chapter, we first delve into traditional sound design practices by observing the most representative arrangements of sonic ‘containers’ and then investigate examples that break the mold, that disrupt pre-established mental constructs to add a layer of thematic complexity. Sonic ‘Containers’ in Traditional Sound Design

Traditional sound design presents clearly defined psycho-cognitive dimensions—that is, whether sounds belong to either the music, dialogue, or sound effects, and whether they emanate from within or outside of the diegesis.1 Each of these dimensions, diagrammed in Figure 2.1, subliminally activates in us the CONTAINER schema.2 Over time, the harmonious orchestration of sonic containers has resulted in listeners constructing stable schemas—arrangements wherein each sonic container encapsulates a distinct psycho-cognitive dimension of the film’s storyworld.3 Most films rely on a basic configuration that includes diegetic voice, diegetic sound effects, and either diegetic or nondiegetic music.4 For example, a scene in Titanic follows a normative paradigm, presenting

DOI: 10.4324/9780429504457-3

24 CONTAINER Schema

FIGURE 2.1

Cognitive dimensions of a soundtrack as ‘containers’.

all three diegetic containers, as shown in Figure 2.2. The diegetic music in the scene— emphasized through the characteristic ‘source-identifying’ shot—contributes toward constructing and further defining the characters’ identities. The sound design of a scene in Star Wars: The Return of the Jedi, diagrammed in Figure 2.3, offers a different, yet still typical, configuration of sonic containers. While in the example from Titanic several source-identifying shots indicate that the music belongs to the diegesis, the absence of such shots in this scene from Star Wars gives the music a different purpose. Here, the music acts as an off-screen narrator informing us about plot developments, with Darth Vader’s leitmotif signaling his imminent presence.5

FIGURE 2.2

Titanic. Normative arrangement of sonic containers. [00:37:30]

CONTAINER Schema

FIGURE 2.3

25

Star Wars: The Return of the Jedi. Normative arrangement of sonic containers. [01:59:45]

Un-Contained Soundtracks

In the context of a clear diegetic and non-diegetic polarity, and a clear distinction between the voice, music, and sound effects, any anomaly or disruption of such normative sound design becomes a marked event, one that prompts us to construct metaphorical interpretations. The following sections unpack unusual interactions among sonic containers, interactions realized in the sound design via three techniques that I call overlap, replacement, and transference.6 Overlap

Overlap entails the simultaneous layering of various sonic containers. The overlap of music, voice, and sound effects, each in a single diegetic form, corresponds to normative sound design. However, the overlap of the diegetic and non-diegetic forms of any single dimension—e.g., the overlap of diegetic and non-diegetic music—denotes a structural disruption that indicates a marked event in the narrative.

FIGURE 2.4

Farewell, My Lovely. Overlap during Philip’s psychedelic experience. [00:50:15]

26 CONTAINER Schema

FIGURE 2.5

The Conversation. Overlap highlighting the tension between Harry’s two worlds. [01:52:00]

In Farewell, My Lovely, a gang captures Philip and takes him to a clandestine brothel for interrogation. He is tortured in the kitchen but resists providing answers to the madam; as a last resort to get information, she injects him full of hallucinatory drugs. The soundtrack begins to blur the boundaries between sonic containers: diegetically ambiguous laughter and clashing of kitchen pots and pans overlap with non-diegetic synthesizer drones and extended instrumental techniques. This disorienting overlap, diagrammed in Figure 2.4, impairs our embodiment of clear sonic containers by dissolving their boundaries and intermingling their contents. As a result, this overlap metaphorically depicts Philip’s mental state and prompts us to embody his psychedelic experience—although trying to maintain his sanity, he is unable to navigate his way through the sonic environment. The soundtrack to The Conversation uses instrumentation to delineate the protagonist’s inner worlds: the diegetic saxophone represents Harry’s need for social interaction, while the non-diegetic piano represents his solitary existence. In a scene toward the end of the film, the overlap between the diegetic saxophone and the non-diegetic piano reflects Harry’s emotional state; as shown in Figure 2.5, these two distinct forms of the music dimension overlap but resist blending. Additionally, by omitting the voice and the sound effects (in either form of diegesis), the sound design directs our attention to the tension emerging from this conflicting overlap in the music, a tension that reflects the dissociation of Harry’s private reality and social life. In Punch Drunk Love, distinct sonic environments delineate the worlds of each protagonist: a non-diegetic waltz represents Lena’s steady and graceful personality, while random and chaotic sounds of a portable harmonium, presented at times diegetically and at times non-diegetically, represent Barry’s unstable and volatile personality. Throughout the film, the soundtrack keeps these incompatible sonic environments separate until the film’s last scene. Lena serenely walks into Barry’s workshop, her gait pairs with the beat of her gentle nondiegetic waltz; Barry sits at the harmonium playing a tranquil melody, learning to master the instrument just as he learns to contain his scattered thoughts. As Lena approaches Barry and lovingly folds her arms around him, the non-diegetic waltz and the diegetic harmonium melody overlap, complementing and embracing each other. Figure 2.6 illustrates this harmonious overlap.

CONTAINER Schema

FIGURE 2.6

27

Punch Drunk Love. Overlap suggesting the characters complementing each other. [01:27:30]

Replacement

Replacement entails omitting sounds we expect in a sonic container and presenting them, in a substitute form, in another container. When the original sound and its substitute share acoustic characteristics, replacement goes almost unnoticed—for instance, omitting the diegetic sound of thunder while including a non-diegetic musical rendition of thunder using percussion instruments. On the other hand, when introducing substitute sounds that deviate significantly from our expectations, replacement calls attention to itself. In such replacements, the substitute sounds’ connotations come to the foreground, effectively imbuing the narrative with subtextual commentary.7 By the 1930s, films had introduced sound. Alexander Nevsky (1938), Eisenstein’s first sound film, circumvented some of sound design’s practical challenges by shooting scenes to music. This resulted in one of the earliest instances of replacement.8 At the climax of the Battle on Ice, Alexander challenges the Teutonic Grand Master to a duel. The two knights duel on horseback while fighters from both camps gather around and watch the brutal exchange— clashing swords, skidding horses. The music maps the action, introducing a furious contest of their respective themes and violent anvil strikes. As diagrammed in Figure 2.7, Prokofiev’s non-diegetic music effectively stands in for the diegetic sound effects; yet, because of the

FIGURE 2.7

Alexander Nevsky. Replacement during a battle. [01:16:25]

28 CONTAINER Schema

FIGURE 2.8

City Lights. Replacement highlighting the absence of content. [00:01:30]

similarity between the sound of clashing swords and a metallic idiophone, this replacement may go unnoticed. Nevertheless, embedding the diegetic sound effects within the nondiegetic music delivers an artful sonic construction that contributes to the viewers’ perception of a stylized battle.9 City Lights (1931), another early film from the sound era, contains music and sound effects, but no voice. Throughout the film, the marked differences between the non-diegetic music and the diegetic voice it often replaces enable the director to embed narrative overtones.10 The film begins with the unveiling of a monument, where dignitaries address the gathered citizens. As the City Mayor begins his speech, a quacking kazoo sound replaces his voice. Moments later, a civic leader approaches the microphone and begins her speech, sounding out a similar garble and quacking, only within a higher, more feminine-sounding register.11 Figure 2.8 presents a diagram of this politically charged replacement. The narrative indeterminacy of the scene—at the beginning of the film and with no setup—leaves ample room for interpretation: since the rhythm and intonation of the kazoo sounds reflect so truthfully the rhythm and intonation of a political speech, this replacement suggests the tendency of political speech to be absent of intelligible and meaningful content.12 While some listeners may argue that the kazoo sounds in City Lights reside at the boundary between music and sound effects, there is absolutely no such ambiguity in The Errand

FIGURE 2.9

The Errand Boy. Replacement during Morty’s impersonating the Chairman. [01:25: 20]

CONTAINER Schema

FIGURE 2.10

29

The Hours. Replacement to connect characters. [01:11:30]

Boy. When Morty, the errand boy at an entertainment company, enters the empty office of the Chairman of the Board, he seizes on the opportunity to sit at the Chairman’s desk and fantasize about being the Chairman. As he addresses imaginary board members, the nondiegetic version of “Blues in Hoss’ Flat” from Count Basie’s Chairman of the Board album replaces Morty’s diegetic voice. This amusing replacement, shown in Figure 2.9, maximizes the comedic power of Morty’s impression by enhancing his already hyperbolic appropriation of the Chairman’s physical gestures. An example from The Hours illustrates the opposite replacement: non-diegetic music replacing diegetic voice. The film explores multiple facets of one personality, a personality that cannot be portrayed by or contained within just one character; instead, all characters in the film contribute to defining a single, abstract persona. At the Hogarth House in 1920s England, Virginia stands alone in the hall as Vanessa, obviously upset, rushes with her daughter, Angelica, to the waiting taxi. In 2001 New York, Clarissa stands in the middle of her living room as Louis leaves, relieved to be on his way. Virginia and Clarissa, each within their own time and place, rest on chairs, still absorbed in the unpleasant departures. The non-diegetic minimalist figures in the underscoring create an almost arithmetical anticipation of a musical cadence. The final cadence, however, is not provided by the music, but by a diegetic sound that crosses through the narratives: the characters’ exhalations. Figure 2.10 illustrates this breathtaking replacement, where the diegetic voice interacts with and ultimately replaces the non-diegetic music, helping construct a shared narrative that crosses time and space. In a scene from Mission Impossible II, a skillfully concealed replacement contributes to character construction. The sound of castanets and animal-like cries of Flamenco dancers resonate through the Andalusian night. It is a large private party, with dancers performing on a raised wooden platform. Through the swirling skirts and pounding heels, Ethan, leader of the Impossible Missions Force, catches a glimpse of Nyah, a highly capable professional thief. She suddenly vanishes from sight, rushing up the stairs, masking her footsteps with the sound of the Flamenco dancers’ steps. In the soundtrack, the diegetic music also functions as diegetic sound effects.13 This replacement, diagrammed in Figure 2.11, contributes to defining Nyah’s quick-witted persona while establishing one of the primary metaphors in the film: the equation of intricate fighting and theft with dance.

30 CONTAINER Schema

FIGURE 2.11

Mission Impossible II. Replacement to cover Nyah’s sounds. [00:12:30]

FIGURE 2.12

Titus. Replacement emphasizing Lavinia’s inability to speak. [01:06:40]

In a scene from Titus, a moving replacement delivers a paralyzing subtext. Lavinia stands amidst a burned field. Her hands and tongue have been cut off to keep her from revealing what she saw. Marcus arrives at the horrific scene and implores: Speak, gentle niece. What stern ungentle hands have lopp’d and hew’d and made thy body bare of her two branches, those sweet ornaments, whose circling shadows kings have sought to sleep in, and might not gain so great a happiness as have thy love? Why dost not speak to me? As Lavinia opens her bloodied mouth and screams, her diegetic voice is omitted and replaced by non-diegetic music.14 This mesmerizing replacement, shown in Figure 2.12, allows the music to take control of the narrative by using a muted flute to portray Lavinia’s inability to speak—her voice, silenced. Transference

Transference entails presenting a sonic event within a specific container and subsequently shifting it to another container.15 Transferences between diegetic music and non-diegetic music containers are quite common and often go unnoticed. Their very presence, however, suggests potential narrative entailments.

CONTAINER Schema

FIGURE 2.13

31

The Milk of Sorrow. Transference depicting a transactional moment. [00:56:00]

The Milk of Sorrow paints a troubling picture of modern Peruvian society, vividly depicting the stark differences between social sectors and their means for negotiating material and spiritual wealth. Fausta, a Peruvian-indigenous woman who possesses a transcendent and spiritual gift for song, works as a maid for Aída, a white upper-class pianist and composer searching for fresh musical materials. Fausta is reluctant to give away her songs, yet Aída promises the financial remuneration that Fausta direly needs. Ultimately, Fausta engages in the painful exchange. In a scene, while Fausta walks silently toward Aída, the non-diegetic singing of Fausta sounds in the background; as Fausta faces Aída, she begins to sing the song aloud, transferring the music to the diegesis. Aída succeeded in drawing Fausta’s most precious song out of her. This transference, shown in Figure 2.13, metaphorically symbolizes the painful passage that Fausta resisted, from spiritual song to material property. Transferences in Inception reveal hidden narrative processes at work, helping the audience navigate their way through a labyrinth of dream stages. In the film, characters enter and leave other characters’ dreams as part of a plan to steal information. The seemingly nondiegetic music heard during characters’ dreams directly relates to the diegetic music heard when characters awaken. According to the film’s narrative, “In a dream, the mind functions more quickly; therefore, time seems to feel more slowly {sic}.” Hence, music in the waking world enters a character’s dream albeit transformed relative to the character’s temporal

FIGURE 2.14

Inception. Transference as Ariadne awakens. [00:27:40]

32 CONTAINER Schema

FIGURE 2.15

The Truman Show. Transference revealing a fabricated reality. [01:09:00]

perception within their dream stage. On a crowded Parisian street, Ariadne and Cobb sit at a café. Ariadne, unsuspecting, looks at the passersby. Cobb suggests, “Our dreams feel real while we’re in them. It’s only when we wake up we realize things were strange,” and asks, “How did we end up here?” Ariadne looks around, confused, unable to remember, and begins to realize she is dreaming. As the cityscape and the dream itself disintegrate, the soundtrack introduces a faint rumble that quickly grows into non-diegetic, elongated, low-sounding ‘BRAAAM . . . BRAAAM’. As her dream collapses, she wakes up in the workshop, “Non, Je Ne Regrette Rien” playing diegetically in the background. A sensitive listener will retrospectively recognize that the initial non-diegetic music contains a version of the song slowed four-fold; this is particularly noticeable because of the corresponding changes in timbre and register—the song’s original lively accompaniment in the trumpet and viola emerges as elongated ‘BRAAAMs’ in a low brass-like sounding instrument within her dream. Ariadne’s first dream thus functions as a key moment in the film—via this awakening transference from what seems to be non-diegetic music to diegetic music, shown in Figure 2.14, we learn (perhaps subliminally) to find our way through the protagonists’ dreams.16 A transference from non-diegetic music to diegetic music frequently denotes a character’s power and control over the narrative. In The Truman Show, Cristof, the manipulative television producer and creator of the show, comes across as a God-like, omnipresent agent, continually observing and controlling Truman’s entire existence. In the scene that opens this chapter, Philip Glass—composer for both the film and the show—appears at the piano, triggering a transference from non-diegetic to diegetic music. This enacted transference between sonic containers, with the Glass cameo, opens a gateway to transcend or even escape the storyworld—it establishes a parallel between the calculated control exercised by Cristof in directing the show and the control exercised by all film directors in constructing an immersive yet fabricated reality. The Father reveals Anthony’s puzzlement as he copes with advanced dementia. All cinematic elements come together to force viewers to experience Anthony’s subjective perspective, embodying his confusion by transgressing aural and (chrono)logical boundaries—actors are not bound to one character and characters are played by multiple actors, dialogues and scenes repeat within an unbroken continuum, décor and lightning morph within single scenes and across scenes. The film opens with Anne rushing through one of London’s

CONTAINER Schema

FIGURE 2.16

33

The Father. Transference to foreground Anthony’s storyworld. [00:01:05]

affluent neighborhoods. Henry Purcell’s opera King Arthur, or the British Worthy, imposes its pace onto the soundtrack, with credit lines and visual cuts metronomically marking the pulse. Impatient, she fetches the keys, opens the door, and calls out, “Dad?” Her anxiety mounts as she checks every room. Anthony is in the studio, peacefully sitting in his armchair, wearing headphones. Surprised to see her, he takes off his headphones—the music stops as if it has been coming through his headphones, transferring the (initially) non-diegetic musical sounds to the diegesis. Via this subtle sound-design manipulation, the soundtrack immerses us within Anthony’s unstable, unreliable reality. As the film further unfolds, subsequent sound-design manipulations will represent his storyworld, while reminding us of the inevitable impermanence of our own reality. Transferences between sonic elements belonging to different forms of diegesis and different layers of a soundtrack (i.e., music, voice, or sound effects) are drastic but effective. In the musical Dancer in the Dark, this sonic manipulation sheds light on the narrative. Selma heads home. She has a degenerative disease that makes her blind, so she uses the railroad tracks as a guiding path. Jeff, a coworker who is romantically interested in Selma, catches up to her and offers her a lift. She smiles and continues walking. Selma hears a freight train in the distance and urges Jeff to stay off the tracks. As the train goes by, Jeff notices that Selma

FIGURE 2.17

Dancer in the Dark. Transference during Selma’s dreamlike drifting. [00:55:10]

34 CONTAINER Schema

FIGURE 2.18

The Conversation. Transference suggesting a sound continues in Harry’s mind. [01:35:50]

is in doubt of where to look for him—she cannot hear him in the noise. He asks, “You can’t see, can you?”, to which she counters, “What is there to see?” Her rhetorical question triggers a transference in the sound design, from the diegetic sound effects to the diegetic and non-diegetic music containers. Selma begins to sing “I Have Seen It All.” Here, the music coexists in both the diegetic and non-diegetic dimensions, illustrating an overlap of sonic containers characteristic of musicals—the singing and the percussive accompaniment of the train sounds are part of the elements seen on screen, but the orchestral accompaniment is not. This wandering transference, shown in Figure 2.17, is vital to our understanding of the narrative: Selma, victimized by her abusive friends and coworkers because of her blindness, is aware of every sound surrounding her, which, in her mind, becomes music. This dreamlike drifting allows Selma to add a veneer of beautiful sonic fantasy onto the tragic reality of her life.17 In another scene from The Conversation, a deafening transference intrudes on the protagonist’s (and our) senses. Harry checks in at a motel, in the room next to where Ann and her coworker are staying. He must listen to their conversations to gain some clarity of their circumstances. Pressing his ear against the wall, he only hears the faintest suggestion of voices coming from their room. In a state of auditory confusion, Harry steps out to the balcony and suddenly sees a bloodied hand on a semi-transparent glass and hears Ann’s scream. In the soundtrack, her scream (diegetic voice container) instantaneously becomes a synthesized sound that resembles a scream (non-diegetic music container). Harry tries to escape reality by turning on the TV and covering his ears with both hands, but the scream (or rather, the impression of a scream) continues to resonate in his thoughts as the non-diegetic sound element persists. As a result, while the film’s plot portrays the transgression of private (sonic) space, the soundtrack maps this transgression onto the cognitive boundaries that define the music, the dialogue, and even the diegesis. Coda

A soundtrack is filled with sounds. During a film, we draw on the CONTAINER schema to organize these sounds into well-compartmentalized streams of information, each revealing

CONTAINER Schema

35

different psycho-cognitive dimensions. While most films use easily identifiable, genre-based conventions with clear-cut distinctions between sonic containers, filmmakers occasionally disrupt these norms to supply a metaphorical subtext. In this chapter, we took a deeper look into these sonic containers and their interactions, and constructed a model for broadening our understanding of film music conventions and music’s potential as a narrative resource. Breaking free from standard sound design formulas, un-contained soundtracks open windows of interpretation that offer audiences a glimpse into previously hidden layers of narrative meaning. However, when stepping outside of normative sound design conventions, we recognize a newly emerging set of conventions used to break those rules: the techniques of overlap, replacement, and transference. For filmmakers, these are effective techniques for a broad range of purposes, from infusing comedic undertones to conveying poignant developments in the story. For analysts, these techniques provide us with a sophisticated framework to interpret a soundtrack’s narrative agency. In The Truman Show, the soundtrack complicates our perception of a fictional narrative by disrupting the storyworld’s sonic boundaries with a sound design transference. This structural transgression not only reveals one character’s power and control over that narrative, but it also reveals the title characters’ inability to escape the superimposed narrative, inducing us, the viewers, to also experience a kind of sonic abuse, one that metaphorically resonates with the emotional and psychological abuse the film condemns.

Notes 1. The terms ‘diegetic’ and ‘non-diegetic’ denote sound sources that are ‘on’ or ‘off’ the screen, respectively. Although numerous scholars discuss the subtleties of diegesis, here I leave the notion of diegesis open, thus avoiding the tendency to subdivide containers ad infinitum. 2. Superimposing the CONTAINER schema’s spatial logic onto our conceptualization of a film’s soundtrack allows us to recognize the metaphorical potential of dynamic interactions among sonic containers. The sonic containers within this framework are fluid constructs, defined by the contained sounds in relation to each other and to the narrative diegesis. For an in-depth discussion of the notion of image schemas, see Appendix II in this volume. 3. To establish a clear distinction between these sonic containers, sound designers monitor and manipulate the sound’s characteristics such as reverb, equalization, loudness, and compression levels. For example, diegetic music features a reverb that matches the physical space where the scene is taking place; if heard from a distance, sounds feature stronger low frequencies; if the sound source moves relative to the camera, the sound will pan or change loudness; and when portraying live music performances, sound designers generally include ‘mistakes’ or other idiosyncrasies characteristic of live settings. 4. Heldt (2013) explores different possible levels of narration stemming from the distinction between diegetic and non-diegetic music. 5. Tan et al. (2008) examine the extent to which pairing a scene with the same music either in its diegetic or non-diegetic forms has an effect on the viewers’ perception of the music and on their interpretations of the film’s narrative. 6. While these interactions may potentially involve any sonic container, this chapter focuses exclusively on interactions between the music (diegetic or non-diegetic) and other sonic containers. 7. Omitting all sound elements but the non-diegetic music is not an instance of replacement. This sound design strategy has become ubiquitous in film, particularly as a means to reveal an emotionally charged angle in the narrative, and therefore is not included in this chapter. In Unfaithful, for instance, a woman becomes enthralled in a passionate extramarital affair; as she leaves the apartment of a handsome stranger and takes a taxi, all sounds but the non-diegetic music are omitted. The soundtrack is thus bestowed with a powerful agency, taking complete control of the

36 CONTAINER Schema

8.

9.

10. 11. 12. 13. 14. 15. 16.

17.

narrative; in particular, the absence of diegetic sounds removes all traces of an objective perspective, forcing the listener into a subjective one. Although by the 1930s sound in film was commonplace, the inclusion of multiple tracks (dialogue, music, sound effects) and the synchronization of sound to picture remained challenging. In fact, throughout the history of cinema, sound design manipulations that have pushed the boundaries of normative conventions have been shaped by aesthetic intentions but have remained constrained by the available technology. Like all image schemas, the CONTAINER schema is a dynamic construct, one that changes over time and across communities of viewers, one that is shaped by past as well as new experiences. While its structure is relatively stable, its contents are largely defined by cultural practices. For instance, a 1920s audience’s perception of a film’s soundtrack through the lens of the CONTAINER schema would result in significantly different contents of sonic containers than a contemporary audience. Furthermore, current sound design practices are fashioning sonic containers with increasingly permeable boundaries, including sounds of ambiguous diegesis and sounds undefined as belonging to the music, voice, or sound effects. The present conceptualization of a soundtrack framed within the CONTAINER schema may therefore deviate from future audiences’ perception of a soundtrack. The score for the film was composed by Charlie Chaplin himself. This is the only reference to the actual voice of characters. Considering Chaplin’s aesthetic rejection of talkies during these years, he possibly embedded this replacement to poke fun at talking films. This replacement illustrates that a sound’s source, rather than its characteristics, defines the container to which a sound belongs. This example may be understood as an instance of “duplex perception” (Bregman, 1990), where a sound potentially corresponds to two sources at once, like rain and applause, or crackling of fire or paper. For a similar effect, see Hitchcock’s The 39 Steps, where the sound of a train’s whistle replaces a woman’s scream. Unlike overlap and replacement, transference entails a syntagmatic (rather than paradigmatic) process. The scene’s sonic environment is foreshadowed in the primary musical gesture of the Main Title soundtrack, becoming a leitmotif to signal the characters’ dream stage. Doll (2018) explores how filmmakers can shape our perception of a scene by obscuring the distinction between diegetic and non-diegetic music, particularly within Inception. This shift to a world wherein everything is controlled by the sounds that overwhelm her senses is visually emphasized by subtle changes in the coloration and dance-like movements of the characters on-screen.

3 LINEARITY Schema

In The Stepford Wives (2004), the music replicates and reveals the protagonist’s reactions. Joanna is a wildly successful, highly paid reality television producer, but her dicey new season presentation goes awry. Immediately after the presentation, a TV network executive summons Joanna to discuss what went wrong. Leaning over, delicately yet assertively, she says, “We have shareholders. We can’t let you sink the network. But we wish you only the best.” Joanna begins to process the meaning behind those words while exchanging looks with her supervisor. As her confident smile slowly turns to mystified terror, a female choir in the soundtrack steadily builds from a wordless murmur to a chilling shriek. The music in the scene duplicates Joanna’s emotional reaction and carries us along with her rising psychological turmoil. Film music moves us. It goes up or down; we feel we go up or down. It gets faster or slower; we feel we get faster or slower. It changes; we change. By traversing any one-dimensional continua, such as pitch, tempo, or loudness, music engages the LINEARITY schema. Engaging this schema during a film prompts us to construct metaphorical mappings between the music and other cinematic domains, like the visuals or the characters’ affective states, eliciting psychophysiological responses that inform our interpretations. In this chapter, we move along a continuum between structural and semantic mappings grounded in the LINEARITY schema.1 While in structural mappings the LINEARITY schema emerges by observing correspondences between the music and the visuals, in semantic mappings the schema emerges from correspondences between the music and the characters’ affective states or the film’s narrative.2 Structural Mappings

Film music often elicits the [PITCH FREQUENCY] IS [MOTION IN VERTICAL SPACE] conceptual metaphor, where upward motion in the visuals correlates with increasing pitch frequencies in the music, and downward motion with decreasing pitch frequencies.3 In a scene from a Tom and Jerry cartoon, for example, the music maps Jerry’s fall from a toy airplane with a descending chromatic figure (and a momentary pause as a brassiere-parachute briefly

DOI: 10.4324/9780429504457-4

38 LINEARITY Schema

FIGURE 3.1

Tom and Jerry Greatest Chases, “Yankee Doodle Mouse”. Jerry’s falls from a toy airplane. [00:05:20]

opens). This mapping of physical movement onto the musical space reinforces the cartoon’s extensive use of ‘Mickey-Mousing’, a technique denoting the synchronization of visual and aural information. Many composers adopted the Mickey-Mousing technique, initially used for cartoons, to underscore non-animated films. The music for King Kong (1933), and much of Max Steiner’s output, is a prime example.4 Kong captured Ann and snatched her to its lair—a cave high above a subterranean pool. Jack is on a mission to rescue her. He approaches the cave, hiding, flattening himself into the crevices of the rocks. While Kong is busy fighting a giant meat-eating bird, Jack seizes on the opportunity to free Ann. Kong roars angrily and turns to them. As they take the only escape route, rappelling down the inner edge of the lair using a sturdy vine, Ann clutching tightly onto Jack, a descending musical gesture in the soundtrack mimics their downward motion. As Kong grabs the vine and begins to pull them back up, the music turns to ascending musical gestures, mapping their upward motion. With little recourse, Jack releases the dangling vine, falling alongside Ann to the subterranean pool. The music once more maps their fall by returning to descending musical gestures. In a scene from The Hindenburg, the music suggests the movement of characters or objects not shown on the screen.5 The passengers are boarding the zeppelin, gazing around, entering a world of luxury and refinement. They fan out, some to their cabins, others to go exploring. At his cabin, looking through the window, Boreth regards his wife on the ground—they exchange a long, loving look. At his own cabin, Kessler also gazes at his wife on the ground—she is at the center of the crowd, waving halfheartedly. Both men remain at their windows and begin to see the crowd, and the world, slowly receding. At this moment in the soundtrack, the music introduces intermittently rising figures that culminate in the Hindenburg theme as the ship flies across the night sky. During this portion of the scene,

FIGURE 3.2

King Kong. Jack rescues Ann. [01:15:30]

LINEARITY Schema

FIGURE 3.3

39

The Hindenburg. The zeppelin departs. [00:26:00]

however, we do not see a character or object moving in the vertical plane; instead, the camera ascends, creating a subjective point of view akin to a passenger’s perspective in the ascending ship. By drawing on the [PITCH FREQUENCY] IS [MOTION IN VERTICAL SPACE] conceptual metaphor, the music helps seal a cognitive gap, resulting in an ingenious compositional device, rather than intricate visual effects, to portray the rising of the ship. The main title sequence of Interview with the Vampire offers a haunting, out-of-body perspective. High above the street level, we circle one of San Francisco’s Golden Gate Bridge towers. A hollow open fifth on the strings’ high register rings in the soundtrack as Christmas lights on the Embarcadero Center illuminate the moonless night. As we fly over the Ferry Building and gently descend on the busy Market Street, the music supplies echoes of the Catholic hymn “Libera Me”, its descending lines first materializing in the children’s voices’ high tessitura and gradually slinking into the strings’ low register. The underscoring in this opening scene prompts us to construct evocative interpretations based on kinetic imagery. While at a surface level the overarching downward musical contour maps the perspective of an unseen vampire descending onto San Francisco, at a deeper level and within the context of the film, the music’s contour arguably suggests a theological and spiritual descent. The LINEARITY schema is not limited to mappings of vertical movement onto the musical space. Often, pitch frequency and loudness metaphorically indicate the size of objects or characters—that is, loud musical figures in the low-frequency range represent large objects or characters, and soft musical figures in the high-frequency range represent small objects and characters. In various crucial scenes, Jurassic Park exploits LINEARITY-based correlations for character construction. Early in the film, we visit the park’s nursery—long white tables covered with dinosaur eggs bathed in infrared light. One egg moves, and the shell begins to crack. In the

FIGURE 3.4

Interview with the Vampire. Descent during the main title sequence.

40 LINEARITY Schema

FIGURE 3.5A

Jurassic Park. Hatching of a tiny baby raptor. [00:28:30]

FIGURE 3.5B

Jurassic Park. Ferocious raptors on the hunt. [01:48:30]

soundtrack, tender musical figures in the high register, performed softly by a gentle choir and celesta, underscore this wondrous moment the hatching of a tiny baby raptor. Later in the film, the now-grown raptors become ferocious predators as they break free and go on the hunt. Tim and Lex, the two kids in the film, hide in the kitchen. The raptors get closer. One raptor stands in the doorway, drawing itself up to its full height; the other stomps into the kitchen, its tail knocking pots and pans from the counter. In the soundtrack, menacing musical figures in the low register performed loudly by brass and timpani underscore this frightening moment, the realization that size does matter. Taken together, the music in these two scenes effectively blends two conceptual metaphors based on the LINEARITY schema: [PITCH FREQUENCY] IS [SIZE] and [LOUDNESS] IS [SIZE].6 Mappings of dynamic parameters are not always consistent with mappings of static parameters. For example, while musical figures in the high register represent small objects and those in the low register represent large objects, music representing dynamic transformations (growing or shrinking) exhibits the opposite tendencies. As the next examples suggest, different principles govern mappings of dynamic parameters—when characters or objects change in size, the music for depicting their growth ascends, and the music for depicting their shrinkage descends.7 In Alice in Wonderland, shrinking in size allows Alice to figure out the meaning of growing up. Alice slithers through the rabbit hole and lands in a round hall with many doors and a three-legged glass table at the center, a small key sitting on top. She grabs the key and opens a small door, about two feet high. There is a lovely garden with a fountain on the other side. She tries to fit through the door, but her shoulders get stuck, so she pulls back. Suddenly, a small bottle with the label “DRINK ME” appears on the glass table. She sniffs the contents and recoils, but shrugs and takes a sip. As Alice begins to shrink, disappearing

LINEARITY Schema

FIGURE 3.6

41

Alice in Wonderland. Alice shrinks after drinking a potion. [00:15:30]

within her now-oversized clothes, the music introduces downward string glissandi. She becomes a few inches tall, the right size to pass through the little door and into the magical garden. In Charlie and the Chocolate Factory, we get a taste of the opposite transformation. Violet is determined to excel in all competitive tasks. As the world record holder in gum chewing, she snatches an experimental stick of gum that delivers all daily meals, including tomato soup, roast beef, and blueberry pie. She munches and grinds relentlessly. Uncomfortably concerned, Wonka suggests, “Spit it out!”, but Violet’s competitive mother flouts Wonka’s advice and cheers, “Keep chewing, kiddo! My little girl’s gonna be the first person in the world to have a chewing gum meal!” She begins to taste the different meals. “Tomato soup, I can feel it running down my throat! It’s changing! Roast beef with baked potato! Crispy skin and butter!” As she gets to the dessert, blueberry pie, Violet turns violet and begins to inflate into a giant blueberry. In the soundtrack, musical figures become higher and higher. Wonka, fascinated, steps back and observes Violet’s bizarre transformation. Such paradoxical relationships between dynamic and static parameters are not pervasive; in fact, the reversal in the pitch-size case is potentially a unique phenomenon. For instance, dynamic and static mappings via the [TEMPO] IS [SPEED OF PHYSICAL MOVEMENT] conceptual metaphor do not exhibit such reversal—that is, fast speeds consistently correspond to fast tempi and slow speeds to slow tempi, regardless of whether these mappings depict static or dynamic parameters.8 A scene in The Incredibles nicely illustrates the mapping of (dynamic) physical movement onto the musical space. A bomb on the elevated train tracks explodes and blows away part of the tracks. As a train approaches, heading straight for the chasm, the non-diegetic action music immerses us in the thrill. Mr. Incredible runs toward the oncoming train, hoping to

FIGURE 3.7

Charlie and Chocolate Factory. Violet turns into a giant blueberry. [01:05:15]

42 LINEARITY Schema

FIGURE 3.8

The Incredibles. Mr. Incredible slows the train to a halt. [00:08:00]

intercept it before it derails. He plants himself on the tracks and braces for full impact, knowing the hit will hurt badly. As he miraculously slows the train to a stop, the music’s tempo also slows down to a grinding halt. Crouching Tiger, Hidden Dragon offers an elegiac spectacle of the mystical facets of martial arts. Two female warriors, Jen and Yu, challenge each other at the center of an interior courtyard. Swords and weapons cover the walls. They begin their fight. Yu scoops up every weapon available, but none is any match for Jen’s Green Destiny sword. The battle intensifies with every new weapon Yu draws on, climaxing as she holds a broken blade at Jen’s neck. Through the scene, the increasing speed of drumming—from about 112 to 186 BPM—and the resulting temporal compression of rhythmic figures map the relentless visuals via the [TEMPO] IS [SPEED OF PHYSICAL MOVEMENT] conceptual metaphor, fiercely thrusting the scene toward the battle’s climax. In the last example, the music, along with the camera movement, editing, and other visual parameters, parallels the increasing aggressiveness of the fighters’ movements. This suggests an additional conceptual metaphor, one that extends beyond musical and spatial-kinetic correspondences, one that captures the characters’ psychological states via the [PSYCHOLOGICAL TENSION] IS [TEMPO] conceptual metaphor, where slow tempi correlate with calm states and fast tempi with agitated states. Note that the mapping direction reverses here, no longer originating in the (concrete) visual domain and shaping the (abstract) musical domain; instead, the music takes on the more concrete role, now suggesting developments in the narrative or inner changes in the characters’ psyches. This shift, from music acting as a ‘target’ domain to music acting as a ‘source’ domain, arguably defines the boundary of the Mickey-Mousing technique and takes us into film music’s semantic sphere.

FIGURE 3.9

Crouching Tiger, Hidden Dragon. Jen and Yu challenge each other. [01:34:10]

LINEARITY Schema

FIGURE 3.10

43

Charlie and the Chocolate Factory. Charlie opens his birthday present. [00:21:00]

Semantic Mappings

By drawing on conceptual metaphors that embed semantic mappings, film music becomes an unseen narrator, revealing the characters’ moods or psychological states and providing glimpses into the narrative’s otherwise hidden facets. PSYCHOLOGICAL TENSION is a frequent target domain in such conceptual metaphors, possibly because it manifests itself more clearly within the narrative than other, more elusive, psychological states. In another scene from Charlie and the Chocolate Factory, Charlie is about to open his birthday present. It is a Wonka bar. Charlie’s parents and grandparents gather around him, hopeful he will find one of the precious golden tickets to visit Wonka’s factory inside the candy. He wants to wait to open the candy, but Grandpa Joe complains, “If you add our ages together, we’re three hundred and eighty-one years old. We don’t wait!” Charlie is anxious in anticipation, but his parents comfort him, “Charlie, you mustn’t be too disappointed if you don’t get one . . . Whatever happens, you’ll still have the candy.” He slowly unwraps the Wonka Whipple-Scrumptious Fudgemallow Delight. In the soundtrack, a crescendo in the strings maps the mounting tension via the [PSYCHOLOGICAL TENSION] IS [LOUDNESS] conceptual metaphor.9 With a brisk move, Charlie rips off the last of the wrapping—no golden ticket.10 Multiple musical parameters may function as a source domain to map the characters’ psychological tension, thus projecting multi-pronged conceptual metaphors. In The Verdict, Frank sees one last shot at salvaging his law career by taking on a medical malpractice case. He is not mentally strong enough, so he relies on Laura, an enigmatic romantic companion, for encouragement and reassurance. Frank’s case experiences a setback. Outside the chambers, he tiredly looks at Laura and admits, “We’re going to lose.”

FIGURE 3.11

The Verdict. Laura belittles Frank. [01:12:50]

44 LINEARITY Schema

FIGURE 3.12

The Stepford Wives. Joanna is terminated from the TV network. [00:09:40]

Laura defiantly retorts: You want me to tell you it’s your fault? It probably is . . . You’re like a kid. You’re coming here like it’s Sunday night, and you want me to say that you’ve got a fever, so you don’t have to go to school . . . Listen! The damned case doesn’t start until tomorrow, and already it’s over for you . . . If you want to be a failure, then do it someplace else. Frank hurries out of the room and shuts the door. In this tense scene, the music begins with a single pitch in the middle register, gradually adding all twelve notes of the chromatic scale and expanding onto a higher register, effectively mapping the increasing tension unfolding from the conversation via two conceptual metaphors: [PSYCHOLOGICAL TENSION] IS [DISSONANCE] and [PSYCHOLOGICAL TENSION] IS [PITCH FREQUENCY]. In the scene from The Stepford Wives that opens this chapter, the music illustrates a different combination of conceptual metaphors mapping the protagonist’s psychological tension. As Joanna realizes she is being fired, a sprawling vocal gesture in the music (performed by a choir) portrays her short but intense psychological turmoil: the increase in loudness maps (and heightens) a comparable rise in pitch frequency, combining the effect of the [PSYCHOLOGICAL TENSION] IS [LOUDNESS] and the [PSYCHOLOGICAL TENSION] IS [PITCH FREQUENCY] conceptual metaphors.11 The potential for combining different musical parameters is vast.12 A scene from Soundless illustrates concomitant relationships between (at least) four musical parameters, all functioning as source domains and reinforcing related conceptual metaphors. Victor, a methodical hitman with a reputation for killing without making a sound, takes one last job. He stares steadily through the scope, aiming at his target through a window in a nearby building. Security workers notice Victor and attempt to alert the target to move away from the window. While static visuals portray Victor’s deep state of concentration and focus as

FIGURE 3.13

Soundless. Victor prepares to shoot. [01:04:05]

LINEARITY Schema

FIGURE 3.14

45

Avatar. Jake mobilizes the Na’vi. [02:02:30]

he prepares to shoot, the underscoring draws on many linear parameters (including loudness, pitch frequency, dissonance, and timbral density) to map the anxiety unfolding among the security workers. As a result, in the thrilling moments anticipating Victor’s shooting, the music establishes a concomitant relationship between musical parameters, reinforcing the related [PSYCHOLOGICAL TENSION] IS [LOUDNESS], [PSYCHOLOGICAL TENSION] IS [PITCH FREQUENCY], [PSYCHOLOGICAL TENSION] IS [DISSONANCE], and [PSYCHOLOGICAL TENSION] IS [TIMBRAL DENSITY] conceptual metaphors. Metaphorical mappings between the music and other facets of a film extend beyond depictions of the characters’ psychological tension. By drawing on our tendencies for musical empathy, metaphorical mappings may help create an aura of collective (and ideological) cohesiveness. In Avatar, humanity discovers the distant moon Pandora, where the indigenous humanoids, the Na’vi, live in harmony with nature. Jake Sully, a wounded former Marine, takes on a mission to infiltrate and exploit the indigenous race in exchange for regaining his mobility. As Jake falls in love with a female Na’vi, he changes allegiances and leads the indigenous humanoids to fight for survival. At a ceremony, Jake rallies the Na’vi. In a subdued and defeated tone, he decries, “The Sky People have sent a message that they can take whatever they want, and no one can stop them.” Gradually, his voice fills with passion and fury to assert, “We will send them a message”, encouraging the crowd to defend itself. In a final note of defiance, he prompts everyone to join forces: “Fly now with me, brothers and sisters! Fly! And we will show the Sky People that this is our land!” The entire tribe responds to Jake’s fervent call to arms, their shouts echoing across the forest. The music in the scene maps the speech’s progression and tone, beginning with a subdued melody in the low register of the celli but steadily intensifying, adding instrumental forces, constructing a coalescing sound mass that climaxes by introducing energetic percussion and a vigorous choir. Here, the [IN-GROUP COHESION] IS [TIMBRAL DENSITY] and [AROUSAL] IS [LOUDNESS] metaphors allow the music to reflect the progression in Jake’s speech, inviting us to empathize with the crowd and the ideology of resistance via musical contagion and subvocalization.13 In Cabaret’s arguably most iconic scene, a young blond boy in a rural beer garden sings “Tomorrow Belongs to Me”, an idyllic song in the traditional German folk style about the beauties of nature, youth, and family life. Moments into the song, the camera reveals the young boy’s Nazi uniform and cuts to other individuals rising and joining in the singing. The build from a soothing solo voice to a strident collective choir accompanies a progression in the characters’ facial expressions, from tender and innocent to angry and threatening. The unfolding collective entrainment depicted in the scene, combined with contextual cues and editing, gradually transforms an idyllic ballad into a militant Nazi anthem.14 This scene

46 LINEARITY Schema

FIGURE 3.15

Cabaret. The crowd joins in singing “Tomorrow Belongs to Me”. [01:18:30]

illustrates how metaphorical mappings that rely on musical contagion, entrainment, and subvocalization may lead individuals to identify with the collective, eliciting an initial aesthetic response to the music that invites individuals into the underlying ideology.15 Metaphorical mappings extend beyond events that unfold within single scenes. Via nuanced, longer-range conceptual metaphors, the soundtrack may help unify entire films by depicting gradual changes in the characters or narrative settings. Short Circuit is a film about Number 5, a robot, coming to life. The music supports this narrative, delineating a semantic opposition between the human and the mechanical. Throughout the film, the [DEGREE OF HUMANITY] IS [TIMBRE] conceptual metaphor unfolds, where electronic timbres denote mechanical beings and acoustic timbres denote human beings. While at the start of the film the music features a synthetic timbre lacking natural overtones, toward the middle, as the robot becomes human-like, the music gradually introduces acoustic timbres, imbuing natural overtones within the musical texture, and by the end, a sixty-piece orchestra swells through the soundtrack as Number 5 starts a new life. In The Conversation, the music also establishes a metaphorical correlation that frames the entire film. Harry Caul is a surveillance expert whose job is to listen into other people’s private lives. Throughout the film, the non-diegetic piano score captures Harry’s private reality through changes in timbre. We thus learn to recognize Harry’s psychological state via the [PSYCHOLOGICAL TENSION] IS [TIMBRAL DISTORTION] conceptual metaphor—enabled by the LINEARITY image schema inherent in both domains—where increases in timbral distortion represent increased levels of psychological distress. Early in the film, as Harry calmly walks home after successfully recording the conversation of a young couple in a park, the main theme transpires through smooth and undistorted piano sounds. The conversation Harry recorded hints at a murder. As he begins to worry about becoming involved in a labyrinth of secrecy and murder, the smooth piano sounds become increasingly distorted. Toward the film’s end, Harry realizes his apartment has been tapped—he himself has been

FIGURE 3.16

Short Circuit. Transformation of Number 5. [00:11:15] [01:05:30] [01:34:00]

LINEARITY Schema

FIGURE 3.17

47

The Conversation. Harry’s psychological journey. [00:09:30] [01:18:00] [01:51:20]

subject to audio surveillance. As he frantically searches for the recording device, a highly distorted piano timbre signals his nervous tension. Moments later, as he gives up hope of finding the recording device, the music briefly stops and returns to the undistorted piano sound, reflecting his surrender. Coda

Film music makes us feel like we are moving along with the characters. By engaging the LINEARITY schema, the music acts as one agent in a multidimensional mapping process, prompting us to construct metaphors that highlight localized surface-level events or deeply hidden narrative layers—but even when localized, such mappings may carry great significance. In The Stepford Wives, the husbands pursue a misogynistic ideal of femininity, turning their wives into gorgeous, subservient, complacent female automatons who lack agency, humanity, or the capacity to feel or express emotions. Joanna’s firing marks the beginning of the transformation that others attempt to impose on her. However, the localized musical event revealing her psychological trajectory via the LINEARITY schema, from circumspect to enraged, allows us to recognize in her the very characteristic absent in the soulless android she must resist becoming. Notes 1. These mappings can be expressed via the conceptual metaphor structure, [A] IS [B], wherein both the source [B] and the target [A] exhibit a one-dimensional continuum. Clustering examples into a conceptual metaphor also clarifies the structure of metaphors and foregrounds the embedded directionality: from [B] to [A]. For an in-depth discussion of the notion of image schemas and conceptual metaphors, see Appendix II. 2. Nevertheless, these are not exclusive categories, and examples may be situated along a broader linear dimension, between purely structural and purely semantic. 3. Many experimental studies on the [PITCH] IS [HEIGHT] conceptual metaphor observe a great degree of consistency within (and beyond) Western cultures. Widmann et al. (2004) used a visual priming paradigm and event related responses (ERP) to establish the consistency of the pitch-verticality correlation; they interpret their results from an ecological perspective, noting that “expectations on forthcoming sounds can speed up responding to environmental changes and can, thus, be a basis for successful adaptation” (p. 709). Much research, however, challenges the assumed vertical orientation, with some favoring other spatial relationships such as lateral orientation (e.g., Rusconi et al., 2006; Wühr & Müsseler, 2002), and some highlighting the absence of both an orientation and a one-dimensional structure. Abril (2001), for instance, explores the ability of

48 LINEARITY Schema

4. 5.

6.

7. 8.

9.

10. 11.

12.

bilingual (Spanish and English) children to accurately recognize and label register shifts in music; although not the object of his study, he points out that English terminology draws on the VERTICALITY schema (with terms such as ‘high’ or ‘low’), while Spanish terminology draws on two unrelated schemas extraneous to the VERTICALITY and LINEARITY schemas (with words such as agudo [‘sharp’, ‘penetrating’] for high registers and grave [‘serious’, ‘severe’] for low registers). Similarly, Zbikowski (2007) questions the (perceived) universality of such metaphorical mappings, pointing out instances of cultural variability: “In Bali and Java pitches are conceived not as ‘high’ and ‘low’ but as ‘small’ and ‘large’. Here the conceptual metaphor is PITCH RELATIONSHIPS ARE RELATIONSHIPS OF PHYSICAL SIZE . . . The Suyá of the Amazon basin do not have an extensive vocabulary for describing pitch relationships. When they are described, however, it is in terms of age: pitches are conceived not as ‘high’ and ‘low’ but as ‘young’ and ‘old’. The conceptual metaphor that guides this mapping is PITCH RELATIONSHIPS ARE AGE RELATIONSHIPS” (pp. 67–68). Taken collectively, these studies suggest that language and culture play “a surprisingly active role in the development and organization of image schemas, contributing not only to cross-linguistic variation but also to some universal similarities among image-schematic concepts” (Dewell, 2005, p. 371). Friedmann (2017) investigates how the chromatic gestures in the “climbing motif” underscoring Kong’s climbing the Empire State Building in King Kong shapes the viewer’s perception of the scene. Music cognition scholars have observed a direct link between the sounds produced by physical movement and those evoked by kinetic imagery. For instance, Clarke (2005) notes that “since sounds in the everyday world specify (among other things) the motional characteristics of their sources, it is inevitable that musical sounds will also specify the fictional movements and gestures of the virtual environment which they conjure up” (p. 74). Some scholars argue that the logic behind the pitch-size association is grounded in our evolutionary biology, stemming from “our life-long experience of correlation between an object’s size and the pitch it would produce. In particular, pitch height is correlated across animal species with body size, as larger species tend to produce lower-pitched sounds” (Eitan, 2013, p. 172). Similarly, Huron et al. (2006) draw on an ethological perspective and suggest that composers often “place a melody in a lower register in order to evoke threatening, dominant or aggressive associations” and conversely, they may place a melody in the high register “to evoke more passive, vulnerable or submissive associations”. (p. 176). This phenomenon is further discussed in Chapter 8 in relation to leitmotif construction. Eitan (2017) suggests that such paradoxical relationships arise from a different set of embodied experiences, ones that activate the [MORE] IS [UP] conceptual metaphor. Many musical parameters, other than pitch and pitch relations, embed a one-dimensional structure but lack a spatial orientation; these parameters include tempo (slow to fast), loudness (soft to loud), density (one layer to multiple layers), consonance (consonant to dissonant), and sometimes timbre (undistorted to distorted). The film music metaphors explored in this volume draw on these musical parameters, mapping their inherent LINEARITY onto spatial, kinetic, and affective domains of human experience. Via recurrent interactions with the environment, we establish a direct relationship between distance and loudness. As a result, a sudden increase in loudness arguably triggers a visceral response, alerting us of a potentially dangerous event such as an object rapidly approaching. In tracing this correlation’s evolutionary origin, Granot and Eitan (2011) argue that it stems from our ecological response to an increase in loudness characteristic of looming threats in natural contexts. In the horror and suspense genres, this technique, known as the ‘red herring’, is used to mislead audiences. Joanna contains her emotions and walks calmly toward the elevators. She steps into the elevator and the doors smoothly close. Alone in the elevator, she releases a blood-curdling scream, one that harkens back to the non-diegetic chilling shriek, almost like enacting agency over the soundtrack by effecting a transference from non-diegetic music to diegetic voice. This very agency is a trademark that will put her apart from the Stepford wives. Arguably, the heightened effect of concomitant relationships has an evolutionary origin. In particular, concomitant relationships in which various parameters intensify (e.g., faster, louder) are environmentally significant for an organism and hence more salient, as these may “imply the approach of a potentially harmful object, raising an organism’s attention and alertness” (Küssner

LINEARITY Schema

49

et al., 2014, p. 12). For instance, Eitan and Granot (2006) note that “the musical dimensions of loudness, pitch, and tempo seem to interact via concomitant intensity levels or contours . . . a pitch rise, a crescendo, and an accelerando are commonly considered intensifying” (p. 225). 13. Contagion, subvocalization, and entrainment are discussed in Chapter 1 as part of a broader mechanism of musical empathy. 14. Belletto (2008) notes that the song “mask[s] a dissonance between image and content, between surface and depth”, and “whatever beauty it may possess, the song’s real function is to consolidate the crowd and marshal them into one uniform voice” (p. 613). In fact, this song is now pervasively (and perversely) used within right-wing neo-Nazi circles for precisely this function. 15. Collective entrainment, as a mode of inducing a collective consciousness distanced form a sense of the individual, manifests itself the most distinctly during religious congregations, patriotic assemblies, social rallies, and sport-related gatherings, which incite individuals into coordinated affect and ideological affiliation.

4 SOURCE-PATH-GOAL & CONTAINER Schemas

In Citizen Kane, the music guides us through the protagonist’s deteriorating marriage. At the Kane residence, Emily sits at the center of the table in the breakfast room. Charles enters, kisses her on the forehead, and sits close. Charles avows, “You are beautiful”, and she flirts back, “Oh, I can’t be.” In the next tableau, some years have passed. Emily, now sitting toward the end of the table, inquires, “Charles, do you know how long you kept me waiting last night while you went to the newspaper for ten minutes? What do you do in a newspaper in the middle of the night?”, to which Charles retorts, “Emily, my dear, your only correspondent is the Inquirer.” More time has passed, and Charles looks older. Emily fires a shot, “Sometimes I think I’d prefer a rival of flesh and blood”, but Charles dismisses her by countering, “Oh, Emily, I don’t spend that much time on the newspaper.” After numerous changes in time, Charles and Emily sit at different ends of the breakfast table, retreating to unbearable silence. The music to the first tableau presents a theme that lacks resolution, opening the space for subsequent tableaux, each presenting a variation on the original theme but failing to provide a conclusive point of repose. As a result, this ‘theme and variations’ design offers well-contained snapshots of their relationship; the inconclusive quality of the music for each tableau, however, drives the narrative forward in time and reveals their relationship’s trajectory, from love to indifference. Film music takes us along the characters’ journeys, through indefinite narrative times and places. Sometimes it takes us to the end of the road, and sometimes it stops just short of the destination, leaving us hanging in suspense and eager for what comes next. In this chapter, we embark on a journey to a metaphorical space where two schemas merge—SOURCEPATH-GOAL and CONTAINER—a space wherein film music reaches a syntactically nuanced narrative potential. Before reaching our ultimate goal, however, we take a detour with two exploratory excursions that offer us a glimpse into how these schemas work independently to guide our responses and interpretations. SOURCE-PATH-GOAL and Narrative Syntax

In our everyday lives, we often resort to metaphors grounded in the SOURCE-PATH-GOAL (SPG) schema to conceptualize the abstract notion of ‘time’ in terms of the more concrete notion of ‘space’ (“Saturday seems so far away”, “The past lies behind us”). Similarly, filmmakers DOI: 10.4324/9780429504457-5

SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.1

51

Film Stars Don’t Die in Liverpool. Passage of time. [00:12:30]

frequently draw on the SPG schema’s spatial logic to suggest the passage and direction of narrative time by carefully controlling camera movement and visual effects (e.g., zooming, tracking, panning), equating space-points (here, there) with narrative time-points (present, future). By framing narrative events as journeys, the cinematography evokes the [PASSAGE OF TIME] IS [MOTION IN SPACE] conceptual metaphor.1 In this first detour, I identify a complementary metaphor that relies on the SPG schema to reveal the passage of time, but one that draws on the music to do so: [DIRECTION OF TIME] IS [DIRECTION OF A SOUND’S ENVELOPE].2 Film Stars Don’t Die in Liverpool draws on both the visuals and the music to guide us through a series of flashbacks. This biographical romance drama recounts the final days of Gloria Grahame and her infatuation with Peter Turner. Peter, a young and up-and-coming actor, sits at the bedside looking after Gloria, a once-glamorous but now ailing actress. As Peter leaves the room, a tracking shot follows him. Soon, the camera breaks free and follows his gaze toward the end of a hallway, leaving him behind and moving swiftly toward the depth-of-field. As the camera halts its motion, it reveals a much younger Gloria, vigorously dancing and exercising. We recognize that the object of Peter’s gaze was the room where he met Gloria years ago and that the camera movement transported us from the present to that past via the [PASSAGE OF TIME] IS [MOTION IN SPACE] conceptual metaphor. This visual metaphor is supported and further clarified by a parallel musical metaphor, the [DIRECTION OF TIME] IS [DIRECTION OF A SOUND’S ENVELOPE] conceptual metaphor. At the beginning of the scene, the music presents piano sounds with their characteristic envelope—a percussive onset or attack (SOURCE), an early decay followed by a more sustained fade (PATH), ultimately reaching silence upon the release of the piano key (GOAL). As the camera moves through the hallway, metaphorically going back in time, the envelope of piano sounds plays backward—from its release (former GOAL) to its attack (former SOURCE), traversing a steady increase of sound (sonic PATH). This sonic manipulation, which results in the retrograde envelope of piano sounds, suggests the backward flow of time by establishing the [DIRECTION OF TIME] IS [DIRECTION OF A SOUND’S ENVELOPE] conceptual metaphor. With this metaphor, the music invites us to suspend our perception of ‘natural time’ and navigate our way through a fictional, constructed ‘narrative time’. Both the visual and the musical metaphors are grounded in the SPG schema, but while the visual metaphor draws on our concrete experiences of moving in physical space, the musical metaphor draws on our ideation of moving through a sound’s envelope. Moulin Rouge! presents a more audibly foregrounded example. The film tells the story of Christian, a young English novelist, who falls for a beautiful singer at the Moulin Rouge. Christian begins writing his memoir. As he reflects on the past, his typing “I first came to Paris one year ago” triggers a tracking shot that briskly moves from his typewriter, through a dance

52 SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.2

Moulin Rouge! Passage of time. [00:04:50]

floor at the Moulin Rouge, to a panoramic view of Paris. In the soundtrack, piano sounds featuring a reversed envelope underscore this extreme zoom-out, metaphorically portraying the backward-flowing narrative time. Once the camera settles, Christian’s non-diegetic voice reveals, “It was 1899, the Summer of Love.” At this moment, the past becomes the present, the piano sounds return to their natural envelope, and the narrative unfolds again forward in time. CONTAINER and Narrative Syntax

Our understanding of musical forms (as large-scale syntactic structures) is almost exclusively grounded in the CONTAINER schema, where similarities and contrasts in musical materials define the formal units, or ‘containers’.3 In this second detour, we explore instances where the CONTAINER schema impinges upon our perception of the music’s formal design and, by extension, shapes our interpretation of a scene. In standard music analysis, the content of such formal units is defined by letters that represent similarity or contrast. For example, ‘A’ denotes a single formal unit containing relatively uniform musical materials; ‘A – B’ denotes a segmentation in two formal units, each containing contrasting materials; and ‘A – B – A – C – A’ denotes a segmentation in five formal units, akin to a rondo form, where the middle and outer units contain contrasting musical materials in relation to the other ones. Composers resort to a one-part formal design (‘A’) to unify lengthy film sequences that depict events unfolding at various times and places. A montage in The Shawshank Redemption illustrates this use of musical underscoring. After 40 years in prison, Brooks walks out on parole. He is disoriented—tears stream down his face. Riding the bus for the first time, he fearfully clutches the seat in front of him. The buzzing city terrifies him. He gets a small, old, grim apartment with heavy wooden beams crossing the ceiling, a rickety bed, and a timeworn desk. He also gets a job bagging groceries at the Foodway, which quickly becomes overwhelming. Only feeding pigeons in the park comforts him, as that brings back memories

FIGURE 4.3

CONTAINER schema of one-part, binary, and rondo forms.

SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.4

53

The Shawshank Redemption. Brooks’s journey of despair. [1:01:00]

of his time in prison. His world is shattered. Dressed in his old suit, he finishes knotting his tie, puts his hat on, and places a letter on the old desk. He takes one last look around and steps up onto a chair. Smiling with inner peace, he shifts his weight on the wobbly chair until it goes out from under him, leaving his feet swinging mid-air, freely. The music accompanying this lengthy montage features an unremitting texture of strings and gentle electronics blended with homorhythmic, solo-piano musical phrases leading to long-sustained sonorities that undermine our perception of a metrical structure and project an introspective, contemplative character. By sustaining this mood throughout, avoiding significant changes in any parameter, the music’s one-part form underscores a lengthy narrative unit depicting Brooks’s journey of despair and hopelessness, unifying that which may otherwise appear segmented or scattered.4 In The Hours, lengthy film sequences connect various narrative threads by weaving in a musical fabric that draws on a minimalist compositional style—continuous accompaniment figures, repeating harmonic frameworks, uninterrupted melodic gestures. For instance, in the montage from this film discussed in Chapter 2, the monothematic music underscores events unfolding at different times and places—the Hogarth House in 1920s England and a New York apartment in 2001. By sustaining a single mood through brief musical interruptions, the underscoring unifies this fractured sequence and contributes to constructing a shared narrative in which the yearnings and fears of the characters intertwine to define a more abstract persona, a single identity transcending time and place.5 In montages depicting simultaneous events unfolding in multiple locations, the music frequently outlines a quasi-rondo form—the music becomes associated with a particular location, halts or changes when the narrative shifts elsewhere, and resumes or changes back with the return to the initial location. In The Taking of Pelham One Two Three, a fourminute sequence juxtaposes simultaneous events unfolding in multiple locations. An active,

FIGURE 4.5

The Hours. The characters’ yearnings and fears intertwine. [01:10:00]

54

SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.6

The Taking of Pelham One Two Three. Simultaneous events in multiple locations. [00:49:40]

fast-paced theme underscores employees at the Federal Reserve, opening gray canvas money bags, sorting bills, and collating the money into packets. The music stops abruptly as the narrative shifts to the Gracie Mansion, New York Mayor’s official residence. The mayor is not feeling well; he is getting a shot in the rear end. Deputy Mayor Warren walks into the bedroom to prompt the mayor to make a public statement about the ongoing hijacking of a subway train and insists, “The mayor of the City of New York, trailing by twenty-two points in all the polls, [should care] enough about seventeen citizens in jeopardy to make a personal appearance in their behalf!” The fast-paced music resumes, in medias res, as we return to the action at the Federal Reserve, with tellers sorting bills by denomination and using machines to pack them faster than the eye can see. Then, nearly all musical layers fade out as we trail through the Subway Command Center, where Transit Police Lieutenant Garber awaits the delivery of the ransom. He exchanges strong words with other officials and rhetorically asks, “How long does it take to get that money together? We’ll never make it. The passengers are dead ducks.” Once again, the fast-paced music returns as we see Federal Reserve clerks assembling bills, selecting ten bundles of fifties and five bundles of hundreds, fastening packets with rubber bands. Seconds later, tensely static music sets in as the narrative shifts to the subway car, where hijackers are arguing about their next move. Blue mentions, “They’ve requested more time . . . I didn’t give it to them.” Green asks, “Suppose they can’t make it?”, to which Blue swears, “Then we do what we said we’d do. There’s no other way.” One last time, the fastpaced music reappears as the narrative returns to the Federal Reserve, where the money is now neatly piled together and loaded into canvas bags. Finally, two guards take the bags and hurry down a corridor toward a gate leading to the security elevators, ultimately placing the bags in a police truck. The ransom money is on its way. Throughout this lengthy scene, the music’s formal design alternates units containing contrasting musical materials, mapping the alternation of environments and events occurring simultaneously in the storyworld, yet coming together to form a unified quasi-rondo, ‘A – (silence) – A – B – (silence) – A – C – A’. Such formal design with interruptions and interpolations is a quintessential compositional strategy to suggest the simultaneous unfolding of various events at different locations. While scenes featuring one-part or rondo forms exploit the music’s unifying potential, scenes featuring a well-defined two-part (binary) form often set up a musical contrast that resonates with the broad themes developed in the film. For example, in an early scene from Vier Minuten [Four Minutes], the music illustrates an ‘A – B’ formal design that foreshadows plot developments. Two rugged workmen and an older woman transport a grand piano on a beat-up pickup truck. With its characteristic fast-paced tempo, rough electric timbres, and hectic guitar riffs, the initial hard rock music brings us into the space of the men driving. The older woman, however, appears irritated. Without hesitation, she reaches over and changes

SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.7

55

Vier Minuten. Contrasting cultural environments. [00:03:10]

the radio dial. A delicate classical piece brings us into her space with its gentle piano timbre, slow unfolding melodies, and refined acoustic balance. Here, the music’s ‘A – B’ formal design (hard rock-to-classical) sets up a dialectic that will play out during the film.6 The Fifth Element traverses multiple binaries—good and evil, light and dark, natural and supernatural, human and non-human. The soundtrack to one of its most iconic scenes supports and further expands on these binaries. Cued by the line “It’s showtime” in the dialogue, the music links the simultaneous (yet spatially disconnected) events and furthers the semantic dichotomy of human and non-human embodied by both, intergalactic quasi-human singer Diva Plavalaguna and cyborg Leeloo. The curtain rises. Diva stands in the center of the stage, a star-filled window behind her. She begins with an almost divine performance of “Il Dolce Suono” [“The Sweet Sound”], a vocally challenging aria from Gaetano Donizetti’s opera Lucia di Lammermoor.7 The piece displays Diva’s strong command of a human vocal range, which she traverses with conjunct motion (i.e., using nearby tones) and arch-shaped melodic contours.8 Meanwhile, still held hostage, Leeloo is bound by her human affordances, unable to free herself and fight back an army of Mangalores. As Donizetti’s aria concludes, the sonic environment changes radically, setting off a contrasting section. Here, Eric Serra’s “Diva Dance” demands a departure from the human vocal affordances—the range of the vocal line now more than doubles that of the preceding section and features angular melodic contours and extreme shifts in register. As Diva engages with supernatural agility in an uncanny sonic spectacle denoting alien dexterity and power, Leeloo miraculously regains her superhuman powers to fight the Mangalores, engaging in a physical spectacle of martial skills.9 As Diva concludes her performance, Leeloo takes down the last Mangalore. To a burst of applause, both Diva and Leeloo take a bow. Through this mesmerizing and thrilling scene, we recognize the human qualities of the ‘A’ section and the extra-human qualities of the ‘B’ section via subvocalization. In turn, our embodiment of the music’s binary formal design guides us through the film’s own conception of a human and non-human dichotomy.

FIGURE 4.8

The Fifth Element. Diva’s and Leeloo’s performances. [01:23:40]

56 SOURCE-PATH-GOAL & CONTAINER Schemas

SOURCE-PATH-GOAL and CONTAINER Combined

While so far we have read scenes through the lens of a single schema, we now turn to moments in which multiple schemas are at play, moments in which the music activates complex conceptual structures that only a combination of schemas can capture.10 Here, we explore the melodic, harmonic, and temporal traits of musical cadences, which elicit in listeners strong responses of arrival, stability, and closure, evoking frameworks of space grounded in both the SPG and CONTAINER schemas.11 Embedded within a film score, musical cadences are powerful rhetorical devices that shape our perception and interpretation of a film’s narrative syntax.12 Because different cadential gestures elicit relative degrees of arrival and closure within the music, when embedded within a film, these musical gestures prompt us to construct interpretations and judgments of relative degrees of arrival and closure within the narrative.13 While cadences ending in the tonic pitch, tonic chord, or at a downbeat provide the most definite sense of stability, becoming a ‘location’ that provides finality and ‘closure’, cadences ending with other pitches, chords, or on weaker metrical positions confer lesser degrees of finality and closure. As a result, the SPG and CONTAINER schemas’ spatial logics enable metaphorical mappings between musical cadences and the film’s narrative, here gathered under the [NARRATIVE CLOSURE] IS [MUSICAL CLOSURE] conceptual metaphor. Two examples from 15 Minutes illustrate the use of melodic and temporal cadential gestures to convey the [NARRATIVE CLOSURE] IS [MUSICAL CLOSURE] conceptual metaphor. In both examples, the music presents two sub-segments featuring nearly identical musical materials yet each ending with a different cadential gesture—the first inconclusive, the second conclusive. Whereas in the first example cadences unfolding within the music’s temporal domain play a vital role in our judgment of arrival and narrative closure, in the second example, cadences unfolding within the pitch domain illustrate the music’s power to shape the narrative arch of a film sequence. In an early scene, Eddie and Jordy engage in a relentless foot chase in a busy Manhattan neighborhood to catch two criminals, Emil and Oleg. In the music, the thrilling drumming on ethnic instruments closely follows the action. Eddie tries to stay in the lead, but he runs out of breath and the criminals are getting away, so he must attempt to take a shot. He crouches, takes aim at Emil, but suddenly lowers his gun, forfeiting his chance to shoot. The drumming vanishes, swiftly fading away with a rhythmic gesture that thwarts the downbeat’s

FIGURE 4.9

15 Minutes. Street pursuit. (Music by Anthony Marinelli & J. Peter Robinson.) [00:48:15] [00:48:30]

SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.10

57

15 Minutes. Oleg dies. (Music by Anthony Marinelli & J. Peter Robinson.) [01:50:00] [01:50:35]

arrival. Only seconds after, he resumes a shooting position, closes an eye, aims, and takes the shot. Simultaneously, the drumming resumes, culminating in the much-expected hit on the downbeat, slamming the narrative segment to an arresting close. As a result, the interplay between rhythm and meter shapes our experience of the scene, bestowing the music with different degrees of resolution, which we map onto the narrative via the [NARRATIVE CLOSURE] IS [MUSICAL CLOSURE] conceptual metaphor.14 Toward the film’s end, Oleg, whom we know as an aspiring film director, is accidentally shot while recording his own movie. The underscoring to Oleg’s final words features a mournful violin melody. Whereas the visuals depict his death by abruptly halting all physical motion, the music hints at a lack of closure by interrupting the violin melody at the subdominant instead of the tonic. As the music suggests, Oleg is merely simulating his death to secure a dramatic, Hollywood-like ending for his film. Upon capturing a successful melodramatic shot of his own death, Oleg resumes the dialogue, and the music returns. A few seconds later, however, Oleg truly dies. A cadential tonic signals this event, providing the musical closure that elicits a sense of narrative closure. While single melodic or rhythmic lines can elicit strong phenomenological responses of arrival and closure, most musical scores accomplish this effect via chordal or harmonic structures. A scene from The Conversation exemplifies the use of harmonic cadences to shape our perception of a lengthy scene. Harry sits absorbed in his thoughts while traveling on an electric bus. In the soundtrack, a bluesy piano accompaniment instills the murky

FIGURE 4.11

The Conversation. Harry visits his mistress. (Music by David Shire.) [00:20:10] [00:27:25]

58 SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.12

Amadeus. Opening and concluding scenes. (Music by Wolfgang A. Mozart.) [00:00:15] [02:52:25]

scene with a contemplative mood. Harry steps out of the bus, crosses a lonely street, enters a building, and stops at the base of a staircase, hesitating. The music also halts, on the subdominant chord, avoiding both musical and narrative closure. Harry approaches an apartment door and enters quietly. Amy, his sweet and unsophisticated mistress, rises from bed in a faded silk Oriental robe. No music on the soundtrack. They chitchat for a while, but Harry becomes uneasy about personal questions and decides to leave. Amy, mystified, susurrates, “I was happy you came tonight, Harry. My toes were dancing under the covers. But I don’t think I’m going to wait for you anymore.” Harry steps outside and boards the electric bus, now heading home. The bluesy piano accompaniment returns, but this time leads to an authentic cadence, eliciting in us a sense of both musical and narrative closure. In Amadeus, instead of shaping a single scene or sequence of scenes, a set of two cadences shapes the entire film. The film opens with a grandiose orchestral D-minor chord, sounding before the film introduces any visual elements. The frail voice of Salieri calling, “Mozart!” resonates through the nighttime streets of Vienna. Then another chord, an A-dominant-seventh, creates a half cadence, to which Salieri’s frail voice again calls in distress, “Mozart!” Through this visual and aural contrapuntal interplay, the film’s first scene punctuates the opening with an unstable musical gesture halting at a dominant chord. By averting tonal conclusion, the orchestral music launches a grandiose narrative space, announcing the beginning of an extraordinary story.15 By the end of the film, as Salieri concludes his narration with an account of Mozart’s funeral, the music presents a plagal cadence in the original key, not only bestowing a sense of closure but aligning with the characteristic religious connotations of plagal cadences.16 August Rush presents a similar example to Amadeus. Freddie, a young kid, becomes captivated by the surrounding sounds which, in his mind, become music. He arrives in New York searching for his parents and immediately drifts into a world controlled by city sounds—car honks, loud machinery noises, and other street sounds turn into an expanding musical texture, climaxing with an inconclusive dominant chord at a weak hypermetrical position announcing the beginning of the narrative.17 In the film’s concluding scene, Freddie conducts an orchestra in New York’s Central Park. As his voiceover conveys a final message, “Music is all around us, all you have to do is listen”, the music provides the much-expected conclusive arrival at the original key’s tonic.

SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.13

59

August Rush. Freddie gets lost in the city’s cacophony. He conducts a symphony orchestra. (Music by Mark Mancina.) [00:28:40] [01:46:40]

Finding Neverland portrays the intimate relationship between writer J. M. Barrie and the Davis family, who inspired him to write his well-known play Peter Pan, or The Boy Who Wouldn’t Grow Up. The film presents a rare example, one in which none of the music cues, even the final one, ever arrive at a final tonic, thus sustaining a sense of perpetual wonder. As a result, although the film addresses the themes of Peter Pan only tangentially, the underscoring provides a direct link to these themes: never-ending childhood and everlasting innocence. In the montage from Citizen Kane that opens this chapter, a theme and variations formal design underscores the protagonist’s crumbling marriage. The first tableau introduces the happy couple with a tender theme, a seductive waltz in E♭-major that soon departs toward unstable tonal regions, cadencing in the unusual (and inconclusive) lowered submediant (C♭-major), opening the space for the depiction of Charles and Emily’s volatile and disintegrating marriage.18 In the next tableau, the theme morphs metrically to duple meter and becomes spirited and playful; in E-major, this variation feels somewhat offset, removed from the original key, and although it cadences in the temporary tonic, its hypermetric premature ending does not project a sense of closure. The next tableau presents a nervous variation

FIGURE 4.14

Finding Neverland. Three moments in the film. (Music by Jan A. P. Kaczmarek.) [00:08:00] [00:28:20] [01:36:20]

60 SOURCE-PATH-GOAL & CONTAINER Schemas

featuring urgent woodwind tremolos that embed martial resonances; harmonic instability further sets in, as this modulating variation first introduces the distant key area of E-minor but hurriedly leads us to D-major. A feisty variation in the B♭-Phrygian mode, which could be heard as prolonging E♭-minor’s (unstable) dominant, underscores the subsequent tableau; a defiant gesture presented at the beginning of this variation repeats obsessively, never resolving, leading us without interruption to the next tableau. The music now becomes somber, with penetrating long notes, the registral break in the theme’s melody suggesting a dislocated, broken marriage; a half cadence toward the end of this variation leads us to the last tableau, but before we get there, echoes of the Gregorian chant known as Dies Irae (“Day of Wrath”) connote the marriage’s dire condition. In the last tableau, the E♭-major tonal area and triple meter harken back to the original theme: a feeble off-beat E♭-pedal tone in the harp, a fragile trace of their once-happy marriage, pulsates through this last variation and punctuates a tenuous final cadence in E♭-major, suggesting a restored yet elusive stability.

FIGURE 4.15A

Citizen Kane. Theme. Breakfast montage’s first tableau. (Music by Bernard Herrmann.) [00:51:50]

FIGURE 4.15B

Citizen Kane. Variation. Breakfast montage’s second tableau. (Music by Bernard Herrmann.) [00:53:10]

SOURCE-PATH-GOAL & CONTAINER Schemas

61

FIGURE 4.15C

Citizen Kane. Variation. Breakfast montage’s third tableau. (Music by Bernard Herrmann.) [00:53:25]

FIGURE 4.15D

Citizen Kane. Variation. Breakfast montage’s fourth tableau. (Music by Bernard Herrmann.) [00:53:55]

FIGURE 4.15E

Citizen Kane. Variation. Breakfast montage’s fifth tableau. (Music by Bernard Herrmann.) [00:54:10]

62 SOURCE-PATH-GOAL & CONTAINER Schemas

FIGURE 4.15F

Citizen Kane. Variation. Breakfast montage’s sixth (and final) tableau. (Music by Bernard Herrmann) [00:54:15]

Coda

Film music transports us. By way of cross-domain metaphors, it takes us along with the characters in their journeys. Whereas musical metaphors grounded in the SPG schema shape our experiences of a narrative’s chronology, those grounded in the CONTAINER schema shape our experiences of a narrative’s segmentation. When combined in the music, the SPG and the CONTAINER schemas suggest syntactic events that set the space for exponentially more nuanced and layered narrative interpretations. Citizen Kane spans one man’s life, from his youth to the aftermath of his death. By foregrounding both the SPG and CONTAINER schemas, the music of the breakfast montage amplifies other elements of the cinematography, allowing us to perceive both continuous time and discrete events. Through fractured timelines and temporal dislocations, enhanced through the musical score, Citizen Kane effectively explores the impermanence of time, the inaccuracy of memory, and the impossibility of recovering the past. Notes 1. Coëgnarts and Kravanja (2012) offer a well-designed framework to flesh out instances of the [PASSAGE OF TIME] IS [MOTION IN SPACE] conceptual metaphor within the cinematography. See also Coëgnarts (2019); Coëgnarts and Kravanja (2015); Forceville and Jeulink (2011). 2. The concept of ‘envelope’, itself a metaphor, denotes the contour of the sound’s intensity through time. It often features four stages: (1) attack, the onset of the sound; (2) decay, the initial quick fade after attack; (3) sustain, the continuation of the sound; and (4) release, the termination of sound, generally caused by ceasing sound production. 3. Whereas conceptualizing sound design categories in terms of the CONTAINER schema, as outlined in Chapter 2, presupposes a synchronic perception of the constituents of the schema, conceptualizing musical form in terms of the CONTAINER schema, outlined in this chapter, presupposes a diachronic perception of the constituents of the schema. For an in-depth discussion of the notion of image schemas and conceptual metaphors, see Appendix II. 4. Thomas Newman’s musical language for the scene brings to mind Aaron Copland’s open voicings, which feature a profusion of fourths, fifths, and ninths. Although the Dorian mode is most prominent, sporadic chromaticism infuses a degree of tonal and modal ambiguity. Exploring a single musical mood to unify a lengthy scene has become a trademark of composer Thomas Newman; this approach is also prominent in his score for American Beauty. 5. See also the metaphorical use of sound design to achieve this effect, as discussed in Chapter 2.

SOURCE-PATH-GOAL & CONTAINER Schemas

63

6. In addition, two sound design manipulations help outline the narrative arc of the film: first, the woman’s changing of the radio dial triggers a transference in the sound design, from non-diegetic to diegetic music; and second, as the characters reach their destination and the narrative unfolds in spaces far removed from the truck, the classical music continues to resonate, suggesting a reversed transference, from diegetic back to non-diegetic. These two transferences in the sound design establish a complementary metaphor that foreshadows the film’s overall plot: the older woman is in control; she nurtures and appeases rebellious and aggressive individuals with the kinds of music traditionally deemed sophisticated and elegant. 7. Donizetti intended this aria to be accompanied by a glass harmonica; instead, the soundtrack features a flute. The eerie sound of a glass harmonica would have worked against the desire to ground this aria in a human, rather than an alien sonic environment. 8. These features are characteristic of vocal writing, grounding the music in the affordances of a human (well-trained) soprano. Additionally, as customary in operatic passages, prosody (i.e., patterns in the text) allows for a flexible rhythmic profile in the solo voice. 9. Here, the wordless vocals create associations with the sound of a Theremin. The use of a Theremin (and more broadly, of electronically generated timbres) has permeated sci-fi films as a convention to represent alien beings since the 1950s. In a survey of soundtracks to sci-fi films, Schmidt (2010) notes, “there is some suggestion that our brains physically interpret electronic sounds as in some way profoundly artificial in relation to the sounds produced by other instruments . . . Thus, no matter how pleasing it may be to the ear, the electronic may always signify both itself and an anxiety about authenticity, and might have always been pre-destined to be alien” (p. 36). 10. In verbal communication, schema combinations may emerge in single words (such as in the verb ‘insert’, which combines the CONTAINER and SPG schemas to specify both boundary and direction) or in verbal metaphors (such as in “Our project hit a wall”, which combines the BLOCKAGE and SPG schemas). 11. When describing musical events related to cadences, we commonly resort to prose—for example, “After a detour [SPG] to the dominant key area, the cadential arrival [SPG] at the tonic closes [CONTAINER] the recapitulation.” In such metaphorical constructions, the SPG and CONTAINER schemas address different facets of our experience. Whereas the SPG schema evokes the syntagmatic unfolding of musical events, the CONTAINER schema evokes the effect a cadence has on shaping a musical (or narrative) structure—i.e., cadences outline the boundaries of a container defined by music-syntactical structures. 12. Lehman (2013) offers an in-depth exploration of the punctuative function of cadences in Hollywood film scores. 13. When describing film events related to narrative syntax, we commonly resort to descriptors such as ‘non-linear story’, ‘cliffhanger’, ‘circular narrative’, ‘plot twist’, or ‘satisfying close’. Thompson et al. (1994) conducted various experiments to examine this phenomenon in relation to the underscoring. Their results suggest that “the final note and harmonic accompaniment in underscoring [at cadential points] can significantly affect viewers’ sense of closure, or finality, of a filmed event” (p. 23) and that “meter and rhythm play an important role in affecting [the] listener’s sense of musical closure” (p. 24). 14. Additionally, a subtle sound design manipulation—the diegetic sound of gunshots replacing the non-diegetic music—suggests that Eddie is in control of the narrative. 15. Motazedian (2016) explores the viability of long-range tonal organization in film music and includes Amadeus as a case study. 16. A plagal cadence is commonly referred to as the ‘Amen’ cadence. Additionally, it is not coincidental that the opening musical gesture and the concluding cadence are in the same key. 17. A subtle sound design manipulation in the scene functions as a metaphor for character development: a smooth transference from diegetic sound effects to non-diegetic music helps depict Freddie’s engagement with music, projecting him as exceptionally sensitive to sounds and music. In the film’s concluding scene, the soundtrack once more extends the boundaries of the CONTAINER schema in the sound design, featuring a transference from diegetic music to non-diegetic music to convey Freddie’s final message. 18. In retrospect, this last sonority (C♭-major) can be reinterpreted (enharmonically) as the dominant of E-major, the key of the next variation.

5 AFFORDANCES

In Tim Burton’s Batman, the music invites us into a ritualized dance, only to trap us in a battle of mythical proportions. Jack and his band of criminals have been set up. They duck into a chemical supply room—a two-story refinery floor accessible through a network of steel ladders and catwalks, filled with huge containers stamped with “Danger! Highly Toxic”. The police are also on the move, hunting for Jack and his men. Jack’s men open fire. The cops shoot back. Their bullets puncture ducts and pipes that spew gas and chemical sludge. A fast-paced waltz in the soundtrack prompts us to engage in the action, distorting the ensuing clash into an unsettling and lethal dance. Shots resonate loudly as Jack scuttles across the elevated walkways, firing at the police, puncturing more ducts and containers, thereby releasing more poisonous chemicals. He spots an exit at the end of a catwalk, but a caped black shadow descends into his path. Giant wings and a yellow-and-black insignia—Batman. His leitmotif swells in the music, assuring us of his presence. As Batman pulls Jack off the ground, one of Jack’s men points a gun at the commissioner’s head and shouts, “Hold it! Let him go, or I’ll do Gordon.” Batman does not move. The cops do not move. All action stops. The music halts its pulse and introduces static, suspenseful sonorities. Batman slowly releases Jack and stands back. Jack straightens his clothes and smirks, “Nice outfit.” Having spotted his gun a few feet away on the catwalk, Jack quickly grabs it and aims at Batman, who has mysteriously vanished. Still poised, gun in hand, Jack aims for one cop and takes him down with a single shot. The music reintroduces fragments of the wicked waltz. As Batman reappears from the shadows, Jack fires point-blank, but Batman swings his heavy cape, and the bullet ricochets toward Jack’s own face. Jack loses balance, stumbles to the edge of the catwalk, and in an agonizing pirouette, topples over, only managing to grab onto the lowest rung—beneath him is a large container full of dark green sludge. Batman leaps to save Jack, reaching out with his black-gloved hand. In the soundtrack, the violins stretch to a seemingly out-of-reach note and struggle to hold tight. Their eyes meet for a moment. Jack is slipping. He cannot hold on any longer and plunges two stories down into the bubbling toxic waste. Throughout this scene, the music thrusts us into Gotham City’s underworld and allows us to embody the actions of its odd, deviant, freakish characters.

DOI: 10.4324/9780429504457-6

Affordances

65

Music is a powerful instrument, a device, a tool—like a hammer, a ladder, a magnifying glass—which affords us opportunities for bodily engagement and extends our capacities to experience and interact with the world around us. It invites us to fight, dance, play, cooperate, unwind. It connects us to the characters’ bodies and minds, and makes us participants in their actions.1 In this chapter, we trace the origin of music’s affordances to our phenomenological experiences in the physical world and explore instances in which the music— through its meter and tonality—operates in concert with other cinematic dimensions to evoke in us physiological and psychological states that provide insight into the characters’ behaviors and the film’s narrative. Affordances of Meter

Understood through the lens of affordances, meter is not a property isolated within the music itself; instead, meter arises from and mediates in our embodied conceptualization of the music. As we entrain to the music’s temporalities, we perceive its meter not as cyclically organized beats but as a feature that allows us to engage in a broad array of interactions, a feature that affords us dancing, marching, or relaxing to the music.2 Because our bodies are symmetrical, most movements related to physical labor unfold in cycles of duple organization (lifting and lowering, pushing and pulling), which translate into onomatopoeic musical renditions of duple meter in work songs ([1—2]–[1—2]– [1—2] . . .).3 Entraining to these songs’ duple meter affords us the kinds of engagement with work that help ease physical effort and ensure greater efficiency and coordination, particularly in communal tasks.4 In films, therefore, music depicting collective work is often rendered as stylized versions of duple meter.5 In Snow White, the song “Whistle While You Work” uses duple meter for a light-hearted, fanciful depiction of work. While wandering in the woods, Snow White and her animal friends sneak into a cute little house. Surprised, Snow White remarks, “Must be seven little children. And from the look of this table, seven untidy little children.” Dust everywhere, dirty little dishes, cobwebs in the fireplace. The bluebirds whistle a military bugle call, and Snow White begins to sing, “Just whistle while you work and cheerfully together, we can tidy up the place.” To the song’s duple meter, she sweeps dust back and forth, birds lift and lower rugs, and chipmunks scrub the tiny clothes from side to side on a turtle’s shell. The background music in duple meter continues as the camera fades to a mine somewhere in the mountains. Four of the Seven Dwarfs are working in a cavern, chipping away at the walls, their picks clinking rhythmically to the duple meter of the music. They break into song: “We

FIGURE 5.1

Snow White. Snow White and her animal friends tidy the place while four Dwarfs work at a mine. [00:18:20]

66 Affordances

FIGURE 5.2

Pirates of the Caribbean: The Curse of the Black Pearl. Pirates march underwater. [01:48:30]

dig, dig, dig, dig, dig, dig, dig from early morn till night. We dig, dig, dig, dig, dig, dig, dig up everything in sight.” In both vignettes, despite the differences in the characters and their forms of work, the music makes us co-participants of the action—while the lyrics and the visuals describe the very phenomenon under discussion here, our entraining to the music’s duple meter affords us the opportunity to internally simulate the kinds of physical labor depicted on-screen.6 Besides aurally depicting work or physical labor, music in duple meter draws our attention to other common cyclical bodily movements of characters, such as walking or marching.7 In a scene from Pirates of the Caribbean: The Curse of the Black Pearl, the moonlight shines down into the deep blue waters. Suddenly, fish scatter as the underwater currents unveil distant figures. In the soundtrack, a plodding duple meter march emerges. The figures appear more clearly, stomping onto the shadows of a shipwreck—these are the pirates, slowly marching across the ocean floor, turning into skeletons when splashed by moonlight. In this captivating scene, the music’s slow duple meter affords us the opportunity to entrain to a spellbinding sunken march and to subliminally trudge along with the ghostly pirates. Triple meter ([1—2—3]–[1—2—3]–[1—2—3] . . .) presents a stark opposition to duple meter in terms of entrainment and affordances. Because triple meter defies our bodies’ symmetrical nature, it commonly characterizes music associated with activities not related—and even directly opposed—to physical labor (such as waltzes or lullabies). Film music often draws on triple meter to depict such activities and make us co-participants of the on-screen action. Finding Neverland offers glimpses of Sir James Matthew Barrie’s escapes from the troubles of adulthood into the imaginative world of childhood. It is a pleasant afternoon in

FIGURE 5.3

Finding Neverland. J. M. Barrie and Porthos waltz. [00:11:35]

Affordances

FIGURE 5.4

67

Batman. Jack’s dancing and murderous moves. [00:36:40]

Kensington Gardens, but James senses the somber mood of Sylvia, a recently widowed mother, and her boys. To lighten the atmosphere, James grabs Porthos, his shaggy Saint Bernard, and begins dancing. A delightful non-diegetic waltz (in triple meter) transports us to an alternative reality—Porthos becomes a great Russian bear courteously dancing and prancing with James within a large circus ring filled with amusing and peculiar characters. Sylvia and the boys clap in enthusiasm, enjoying the show. The music’s triple meter affords us a journey into J. M. Barrie’s transformative, fantastical, surreal dance, and invites us to “believe, believe, believe!” As suggested in the scene from Finding Neverland, the odd number of impulses in triple meter, which does not align well with our evenly shaped bodies, carries semantic associations that characterize ‘odd’ settings or individuals. In the scene from Batman with which I open the chapter, Jack falls into bubbling chemical waste that radically alters his face, giving him a clown-like appearance. In a later scene, Jack pays a visit to Grissom, the most powerful crime lord in all of Gotham City, the one who set him up. Grissom surreptitiously reaches for a gun. Standing in the shadows, Jack warns, “Don’t bother. Jack’s dead, my friend. You can call me Joker.” He steps into the light, flings away his hat, and reveals his hideous glory. A non-diegetic carnivalesque waltz assaults the scene as Joker’s Mickey-Moused murderous moves become an odd and deadly dance.3 Subliminally, we entrain to the music’s triple meter, which affords us a glimpse into Joker’s bizarre, anomalous, offbeat world. Asymmetric meters are complex temporal structures that defy organization into cyclic groups of two or three. By not affording us the organic entrainment characteristic of duple or triple meters, asymmetric meters present a physical challenge to both performers and

FIGURE 5.5

Mission Impossible. Ethan accomplishing a near-impossible task. [01:42:30]

68 Affordances

FIGURE 5.6

The Taking of Pelham One Two Three. Bankers struggle to collect the money. [00:49:40]

listeners.8 Quintuple meter, for instance, although not impossible to entrain to, demands considerable effort. The main theme for the Mission Impossible TV series (later used for the films) captures the near-impossibility of the teams’ quests through the complexity of the music’s quintuple meter ([1—2—3—4—5]–[1—2—3—4—5] . . .). During the TV series and the films, the main theme enters at critical moments, in which the protagonists are about to complete their (seemingly) impossible missions. The music’s complex rhythmic structure (a subdivided quintuple pulse organized as 3 + 3+2 + 2) impinges on the film’s kinesthetic dimension—by breaking the divide between the film’s fictional world and our bodily reality, the music in these scenes demands a complex physiological response, one that brings us along with the characters to the limits of our shared affordances. In a scene from The Taking of Pelham One Two Three, also discussed in Chapter 4, city officials struggle against time to collect and count the terrorists’ million-dollar ransom. Moments before the cue begins, Police Lt. Zachary Garber exclaims, “Your instructions were complicated. The money has to be counted, stacked, transported uptown. It just isn’t physically possible!” To Garber’s list of bodily activities suggesting a duple work meter, the music’s 7/4 asymmetric meter ([1—2—3—4—5—6—7]–[1—2—3—4—5—6—7] . . .) lends the scene an unyielding and relentless feel, affording us a visceral understanding of the bankers’ desperate attempt to comply with the unreasonable demands. The more complex the temporal structures, the less we can entrain to them. For most of us, asymmetric meters with a higher number of beats, or phrases with changing meters, are at the boundary of entrainment. This resistance to entrainment instills in us a high degree of uncertainty about the downbeat’s arrival, eliciting an intensely unsettling physical response uniquely suitable for the horror and suspense genres.9 The main theme from the iconic horror film The Exorcist features 15-tuple meters—shifted and subdivided as 3½ + 3½ + 3½ + 3½ + 1,

FIGURE 5.7

Main theme for The Exorcist. (Music by Mike Oldfield.)

Affordances

FIGURE 5.8

69

The Matrix. Trinity soars in the air. [01:30:15]

as shown on Figure 5.7. As a result, the asymmetric meter presented in the main titles supports and advances the aesthetics behind this film, which exploit uncertainty to elicit fear, anxiety, and distress. By not affording us the opportunity to establish a regular pattern of entrainment, such complex asymmetric meters prime us for the terrifying events about to unfold and trigger the adrenaline rush we crave. Nevertheless, although these responses are usually regarded as negative emotions, within the safe environment of a movie’s storyworld, these become, for some, highly pleasurable experiences. Entrainment is a vital component in constructing our temporal reality—our physical time impinges on our psychological time. In the absence of a metric structure of cyclical beat patterns, we experience a stop in motor activity that evokes a postural immobility akin to a resting state, suggesting a chronological stretching of the ‘now’, a static timelessness, a lengthening of our experience of the present. In The Matrix, fleeting moments of suspended action are framed on either side by intense pulsating activity in duple meter. Trinity and the Key Maker are atop a speeding cargo truck carrying motorcycles. Trinity looks back and sees an explosion on the horizon. Link, over the phone, urges her, “Keep moving.” The music prods us with its relentless duple meter. She grabs the Key Maker, “Let’s go”, and they rush to the front of the truck. She straddles the first bike, and the Key Maker climbs on behind her. As she shoots the chain and pops the clutch, the back tire screeches and the motorbike dashes up the final wedge of cargo ramp, leaping into the air, soaring over the truck’s cabin, virtually suspending time into infinity. The ametric music underscoring this moment maps the slow-motion visuals and thwarts our entrainment cycle, affording us the opportunity to soar with Trinity, defying time and space. Entraining to the music’s metrical structure is not just an individualized psycho-physiological phenomenon, but one that reflects collective practices.10 We are constantly immersed in social and cultural environments defined by bodily activities that prompt us to construct associations between motor patterns, the musical metrical structures that afford those patterns, and the socio-cultural environment where such motor patterns unfold. Collective entrainment is one of the primary mechanisms for increasing our individual capacities, enabling critical human activities that rely on social interaction, such as courtship, hunting, and building. Film music often draws on this phenomenon by extending the implications of meter beyond the purely physical realm, using meter as a marker to connote characteristic psycho-social activities—for instance, duple meter for competition or disparagement, triple meter for seduction or adulation. 300 depicts the brutal reality of Spartan soldiers as they engage in the epically fierce Battle of Thermopylae. The non-diegetic voice of Leonidas, the Spartan King, assures his

70 Affordances

FIGURE 5.9

300. Leonidas versus a rhinoceros. [01:10:00]

men, “We can hold the Hot Gates. We can win.” Spartans erupt, “Haawwooo!”, which sets off a slow-motion montage underscored with duple meter music. An armored rhinoceros menacingly approaches the Spartans. Firmly standing in front of his brave men, Leonidas shoots his spear at the beast, his body straight, his muscles unquivering under the heavy armor, only the hem of his tattered crimson cape pushed gently by the wind. Finally, the creature collapses, its horns and thick folded skin burrowing into the muddy ground, growling its last growl at the Spartan King’s feet. Throughout the entire montage, the music does not synchronize with the slow-motion visuals. Instead, the music’s duple meter elicits and reinforces explicit psycho-social facets of the narrative, affording us bodily engagement with the film’s martial associations.11 In Punch Drunk Love, the music colors our interpretations of the visuals and dialogue. On a mission to repair his relationship with Lena, Barry rushes out of the elevator, carrying his harmonium. He careens down the hallway, first going right, then twirling left at the corner, swaying along with the heavy instrument. The music does not synchronize with his running; instead, it introduces a waltzing meter, suggesting a nascent romance. He arrives, puts the instrument on the floor, and rings the bell. As Lena answers, he rambles: Lena, I’m so sorry. I’m so sorry that I left you at the hospital. I called a phone sex line. I called a phone sex line before I met you, and four blond brothers came after me, and they hurt you, and I’m sorry. And then I had to leave because I wanted to make sure you never got hurt again. And I have a lot of pudding, and in six to eight weeks it can be redeemed. So, if you can just give me that much time, I think I can get enough mileage to fly with you wherever you have to go if you have to travel for your work, because I don’t want to be anywhere without you. So, can you just let me redeem the mileage? Throughout Barry’s lengthy and scattered monologue, the characters remain at opposite sides of the visual frame, their contrasting dress colors—Barry in blue, Lena in red— intensifying the dichotomy and polarization.12 The music’s triple meter, however, softens our perception of the confrontation and foreshadows a harmonious resolution. Lena reminds Barry, “You left me at the hospital. You can’t do that.” Barry reiterates, “If you just give me six to eight weeks, I can redeem the mileage, and I can go with you wherever you have to travel.” For a long moment, they gaze at each other in understanding, smile, and ultimately kiss. As the couple embraces and gently moves in unison, we join them in entraining to the triple meter of the romantic non-diegetic music.

Affordances

FIGURE 5.10

71

Punch Drunk Love. Barry rushes to Lena’s apartment. [01:24:20]

The connotative sphere of musical meter extends even further. The music’s affordances, and the kinds of bodily engagement these connote, serve as markers of socio-cultural norms and values. As a result, via entrainment we shape our social identities. The duple versus triple dichotomy, for example, often connotes the poor versus the wealthy, or low versus high social status. Because the music’s affordances are integral to delineating cultural constructs and social boundaries, entraining to the music’s meter allows us a more nuanced understanding of a scene.13 In An Education, David attempts to convince teenage Jenny to accept the questionable ethical acts that allow them to afford their extravagant lifestyle of wealth, refinement, and leisure. David, Jenny, and a couple of friends pull up on a luxurious Bristol just outside their Bedford Square flat. As everyone else heads for a drink, Jenny flees, but David catches up with her on the street. He explains, “It was an old map cooped up in that miserable little cottage, and she didn’t even know what it was. We liberated it.” Nodding reluctantly, Jenny snorts, “Liberated! That’s one word for it.” David counters: Oh, don’t be bourgeois. I know you have fun with us. You drink everything I put in front of you down in one, every last drop, and then you slam your glass down on the bar and ask for more. We’re not clever like you, so we have to be clever in other ways. In the music, ametric string textures quietly underscore Jenny’s inner struggle. She nods reluctantly, but David continues trying to persuade her: If you don’t like it, then I will understand, and you can go back to Twickenham and listen to the Home Service and do your Latin homework. But these weekends, and the restaurants and the concerts don’t grow on trees. This is who we are, Jenny. As David teases her with the prospects of a lavish lifestyle, the melody teases us with echoes of a triple meter. He holds out his hand, inviting her to join them as they are.14 After a few wavering moments, Jenny smiles and takes his hand, consenting to David’s tempting offer. As David pulls her toward him, the musical underscoring swiftly changes to a well-defined triple meter, and both characters begin to dance a (non-diegetic) waltz. Here, the triple meter in the music functions as a commentary on the narrative discourse, signaling Jenny’s rejection of a work ethic and embrace of a life of leisure only possible at a higher social status.

72

Affordances

FIGURE 5.11

An Education. David persuades Jenny. (Music by Paul Englishby.) [00:47:50]

Affordances of Tonality

Tonality extends beyond the hierarchical relationship of tones in a musical system. The perceived stability and instability of particular tones affect our (inter)active behavior with the music—as we embody the music’s tonality, we subliminally attune our motor systems to a broad array of potential interactions. Therefore, understood through the lens of affordances, tonality is not a property isolated within the music itself, but one that mediates between our embodied conceptualization of the music and our interaction with the environment. In this second portion of the chapter, we explore tonal frameworks that allow (or hinder) our perception of a musical gravitational center, tonal frameworks that afford us (or not) the opportunity to feel physically and mentally grounded. Insights on gravitational disturbances outside of music shed light on the logic behind the affordances of gravity in music. For instance, aeronautics experiments that simulate weightlessness conditions characteristic of space offer enthralling insights into the impact of gravitational forces (and their absence) on human physiology, biochemistry, cognition, and behavior.15 Much of this research reminds us that the feeling of disorientation associated with gravitational disturbances stems from ‘cognitive dissonance’—conflicting inter-sensory information reaching our brains, such as our vision telling us we are stationary while our hearing tells us we are moving.16 While many inter-sensory conflicts are possible—emerging from contradictory kinetic, proprioceptive, vestibular, visual, aural, tactile, olfactory, or other sensory information—the visual-vestibular conflict is particularly salient. From the literature on cognitive dissonance transpires the fact that our vestibular systems’ strong connection to our hearing apparatus—the vestibular system is located inside the inner ear—may explain music’s unique ability to elicit disturbances akin to the absence of a gravitational center.17 In films, the music accompanying zero-gravity or extraterrestrial scenes typically draws on symmetrical pitch collections.18 The inherent structure of these collections, wherein all pitches may equally serve as a tonic, contributes to equalizing each pitch’s pulling power by dissolving their primacy to function as a (musical) gravitational center.19 By dissolving tonality, and thereby not affording us the potential to experience tonal grounding, symmetrical constructs contribute to suspending our embodiment of physical gravity, particularly when coupled with equally disorienting visuals.20 The ensuing cognitive dissonance—the music

Affordances

73

telling us there is no gravitational center while our proprioceptive sense tells us otherwise— has therefore implications for our embodied responses to symmetrical pitch collections.21 Peter Hyams’s 2010: The Year We Made Contact, the sequel to Stanley Kubrick’s 2001: A Space Odyssey, chronicles a space mission to one of Jupiter’s seemingly inhospitable moons aboard the Discovery Two spacecraft. The underscoring for scenes showing the Discovery Two floating in space presents symmetrical pitch collections of various cardinalities— number of notes—to portend the kinds of bodily engagement that escape terrestrial forces. The first time we see the spaceship, an ominous tritone—a two-note collection that divides the octave exactly in half—saturates the musical fabric, envelops our senses, and contributes to disturbing our proprioceptive sense of gravitation.22 The second time, a hovering whole-tone

FIGURE 5.12A

First appearance of the Discovery Two spacecraft. (Music by David Shire.)

FIGURE 5.12B

The Discovery Two approaches Europa. (Music by David Shire.)

FIGURE 5.12C

The Discovery Two about to complete its mission. (Music by David Shire.)

FIGURE 5.12D

2010: The Year We Made Contact. Different moments in the Discovery Two’s journey. [00:22:00] [00:32:50] [01:37:35]

74

Affordances

FIGURE 5.13

Alice in Wonderland. Alice falls through the rabbit hole. [00:15:30]

collection—a set of six equally distanced notes—further lifts us and affords us an opportunity to sense the weightlessness depicted in the visuals. As the film unfolds, the underscoring for scenes depicting (or intended to elicit in us the sensation of) zero-gravity becomes denser, drawing on collections of higher cardinalities. For instance, well into the film, the music presents a dodecaphonic collection (gesturally segmented into two transpositions of a hexatonic collection) during a countdown sequence. Here, the music affords us a sense of weightlessness and infuses a sense of suspense and uncertainty that supports the events in the narrative. Scenes unfolding in surreal worlds—not governed by the earthly gravitational forces we recognize—are also commonly underscored with symmetrical pitch collections. In Disney’s animated version of Alice in Wonderland (1951), as Alice falls through a rabbit hole and through the center of the Earth, the objects she encounters behave strangely, not abiding by gravity’s physical laws. A downward scale outlining a whole-tone collection [C, D, E, F#, G#, A#] underscores the initial fall, and as gravity further dissolves throughout her fall, the music introduces an effect akin to a Shepard tone, simulating a continuous descending shift.23 Finally, as Alice enters an unknown world transformed by altered perspectives, the music again features a whole-tone collection, which further contributes to the suspense and uncertainty in the narrative.24 As these examples begin to suggest, the affordances of symmetrical constructs extend beyond our bodily capacities and into our psychological faculties. Because these constructs do not afford us the opportunity to define a clear tonal center, they elicit feelings of suspense, uncertainty, confusion, irresolution, and lack of direction.25 Therefore, in scenes taking place on Earth—where gravity is intact—symmetrical constructs generally serve as psychological markers of instability, affording us a window into the characters in their dream states, hallucinatory episodes, unusual passages of time, or trips to the unknown. Dream worlds are often unfettered by the laws of gravity. Even when firmly grounded on Earth, while dreaming, we may experience proprioceptive hallucinations, such as the sensations of floating, flying, or falling. Symmetrical pitch collections are therefore exceptionally suited for underscoring dream sequences and have been a staple of filmic dream worlds. In the iconic dream sequence from Alfred Hitchcock’s psychoanalytic thriller Spellbound, the music captures John Ballantyne’s physical and psychological lack of grounding. John leans on a comfortable chair and describes his nightmares to a psychiatrist. His recollections depict forces of gravity and motion in space: I can’t make out just what sort of a place it was. He was leaning over the sloping roof of a high building, the man with the beard. I yelled for him to watch out. Then he went over, slowly, with his feet in the air. Then I saw the proprietor again, the man in the mask. He

Affordances

FIGURE 5.14

75

Spellbound. John’s surreal dream. [01:25:00]

was hiding behind a tall chimney, and he had a small wheel in his hand. I saw him drop the wheel on the roof (emphasis added). Despite the physical descriptions, the character is unable to identify the place, perhaps because of the spatial and gravitational oddities the dream projects. Salvador Dalí’s vignettes for the film, featuring free-floating eyes, slanted surfaces, falling objects, and timeless clocks, submerge us in a surreal world governed by contradictory conditions and irrational juxtapositions. The musical fabric, saturated with figures using symmetrical constructs—whole-tone [C, D, E, F#, G#, A#], octatonic [A, B♭ C, D♭, D#, E, F#, G], and dodecaphonic collections—has a dual function: at face value, it contributes to immersing us in the suspended reality of a dream sequence, but ultimately, it is vital in eliciting a pale shade of the protagonist’s mental instability. When music based on such pitch collections does not accompany visuals that portray altered realities or gravitational disturbances, we intuitively draw a parallel to the character’s psyche or to the unfolding narrative instead. Two scenes from The Simpsons TV series illustrate this use of music. In the episode “Them, Robot”, a factory worker enters a supply closet

FIGURE 5.15A

The Simpsons, “Them, Robot”. Worker procures paperclips. [00:01:20]

FIGURE 5.15B

The Simpsons, “Beware My Cheating Bart”. Homer reads a fortune cookie message. [00:01:10]

76

Affordances

FIGURE 5.16

Back to the Future. The car vanishes. [00:18:30]

to procure one standard-size paper clip; as he exits the closet, a green rod falls to the ground and prevents the door from closing. In another episode, “Beware My Cheating Bart”, Homer finishes lunch at a mall, opens a fortune cookie that reads “Eat less, live longer”, and begins to feel dizzy. Although the events of these scenes seem inconsequential, the musical fabric— which draws exclusively on a whole-tone collection [C, D, E, F#, G#, A#]—infuses suspenseful undertones, commenting on these events, foreshadowing a destabilized narrative. Since disorientating events in the narrative may stem from uncertainty about time, composers often introduce symmetrical pitch collections to evoke the lack of chronological direction. In Back to the Future, Marty is accidentally thrown back into the 1950s during a failed experiment by his eccentric scientist friend Emmet “Doc” Brown. Toward the beginning of the film, they are experimenting with time, attempting to send a remote-controlled car to the future with Emmet’s dog in the driver’s seat. The DeLorean roars to life as Emmet manipulates its remote control. It takes off. The speedometer passes 20, then 50, then 80 . . . 85 . . . 88 . . . then bam! A sharp blast of air and electricity hits Emmet and Marty. Marty blinks in disbelief as the car vanishes, only leaving behind a trail of fire and the car’s vanity plate, which reads “OUTATIME.” Marty asks, “Where the hell are they?”, to which Emmet retorts, “The appropriate question is: When the hell are they?” The music throughout the lengthy scene draws on an octatonic collection [G, A♭, A#, B, C#, D, E, F] affording us an opportunity to partake in the characters’ new experience of time, which now lacks the linear chronological forces we recognize. Combined Metrical and Tonal Affordances

The combination of metrical and tonal affordances is exceptionally powerful in guiding our bodily engagement with the music and informing the interpretations we construct based upon such engagement. In fact, the expressive power of combined musical affordances is such that during a film, our bodily engagement and interpretation of a scene may go counter to those suggested by the visuals. In scenes taking place on Earth, for example, the music may suspend our sense of narrative direction by not affording us the ability to locate a clear tonal center or to entrain to a clear metrical framework; alternatively, in the presence of visuals that suggest weightlessness, a clear metrical and tonal framework may ground our experience to simulate earthbound activities. The following examples illustrate the power of combined affordances by drawing on this gravitational phenomenon. The chapter’s opening example from Batman, the shootout scene, illustrates an effective combination of metrical and tonal affordances to generate suspense in a scene taking place on Earth. The initial music accompanying the relentless confrontation between

Affordances

77

the police, Jack’s gang, and Batman, projects a well-defined tonal gravitational center via the C-minor pitch collection. Its polymetrical framework, a quick waltz within an underlying duple hypermetric structure, signals both the current confrontation and foreshadows Jack’s transformation into the Joker. Unexpectedly, as one of Jack’s men shouts, “Hold it!” and threatens to take down the commissioner, the confrontation stops. Will Batman make a move? Will Jack escape the scene? The music supports the stop in the action and the suspense that emerges from it by halting its pulse and introducing a static whole-tone collection [C, D, E, F#, G#, A#]. This drastic change in the music’s affordances affects us viscerally, guiding our embodied engagement with the music and our interpretations.

FIGURE 5.17A

The shootout begins. (Music by Danny Elfman.)

FIGURE 5.17B

The shootout comes to a halt. (Music by Danny Elfman.)

FIGURE 5.17C

Batman. Halt in the shootout. [00:28:10]

78

Affordances

FIGURE 5.18A

Atonal, ametric music in The Matrix. (Music by Don Davis.)

FIGURE 5.18B

The Matrix. Neo on the edge. [00:15:55]

The Matrix is also set on Earth, but the film presents earthly life as an illusion, one that a chosen few can escape by awakening from a dream state. Neo is a chosen one, but he does not yet know it. He is on the run, attempting to escape from ‘agents’—sentient eradication programs within the source code of the matrix, manifested as men in dark green suits and square sunglasses. On his left, there is a window, which he can use to get to the roof. He debates with himself, “It’s insane!”, yet takes his chances, climbs out the window, and begins walking on the narrow ledge of the skyscraper. A close-up of Neo reveals his psychological disorientation. A sudden gust of wind knocks Neo off-balance, forcing him to drop his phone so that he can cling tightly to the building. Startled, he watches as the abyss swallows the phone. The music reinforces this feeling of disorientation via its (a)tonal and (a)metrical design: a near-symmetrical pitch collection does not afford us a firm gravitational point, and seemingly random metrical displacements of a musical figure hinder our entraining to a pulse. Here, the musical and visual devices capture the protagonist’s physical and psychological disorientation, and do more: they foreshadow the revelations to come and evoke the idea that the world Neo lives in may not be governed by our gravitational and chronological realities. In another scene from The Taking of Pelham One Two Three, sixteen passengers and the conductor have been held hostage inside a subway car for hours. Blue, a mercenary, meanders into the subway car and ominously orders, “You. Stand up, please.” The passengers warily turn to look at whomever Blue has indicated. The conductor slowly looks up and, in a broken voice, asks, “You mean me?” Blue quietly leads the conductor by one arm to the rear of the car—all passengers’ eyes are on the conductor as he passes by, his free hand hesitantly clinging to one overhead strap after another. The music accompanying the scene does not project a beat or pulse, and draws on a nine-note chromatic collection to undermine any sense of stability. Like the conductor, we seek a point of support, but the music gives us nothing to grab onto. Instead, the ametric and atonal music maps the static visuals

Affordances

79

FIGURE 5.19A

Nine-note chromatic collection, ametric music in The Taking of Pelham One Two Three. (Music by David Shire.)

FIGURE 5.19B

The Taking of Pelham One Two Three. Hostages wait endlessly in the subway car. [01:05:45]

and effectively underscores the hostages’ unresolved tension, affording us an attunement to their unending expectancy and uncertainty. Often, scenes in outer space illustrate the expressive power of music to counter the visuals by providing the (missing) gravitational pull. Toward the beginning of 2001: A Space Odyssey, astronauts aboard the Discovery One engage in earth-like activities. A space attendant brings food to the astronauts, walking along a circular hallway, carrying a tray with food, ending up upside-down; however, neither she nor the food on the tray ‘fall’ to what we perceive as the gravitational center.26 The music accompanying this montage, Johann Strauss’s “Blue Danube” waltz, counteracts these visual depictions of zero-gravity by establishing a distinct tonal center (D-major) affording us a gravitational center, and a well-defined triple meter affording us earth-like kinds of bodily engagement. Therefore, even when the montage unfolds within zero-gravity environments, the music contributes to recreating the feel of earthbound experiences, reflecting the ‘terrestrial’ sensations of the astronauts.27

FIGURE 5.20

2001: A Space Odyssey. Attendant brings food to astronauts. [00:35:20]

80 Affordances

FIGURE 5.21

WALL-E. WALL-E and EVE dance around the Axiom. [01:00:05]

WALL-E also presents a scene in outer space featuring music that helps reconstruct an earthbound experience. In the distant future, humankind has abandoned Earth. WALL-E, a small garbage-collecting robot, is left behind by himself to clean up the rubble. As he tinkers with the trash he collects, he inadvertently becomes fascinated with Earth’s history, especially show tunes and musicals. EVE, a reconnaissance ‘female’ robot, arrives in search of living organisms. WALL-E falls in love with EVE, and both embark on a fantastic journey through space, hitching a ride on the outside of Axiom, a spacecraft carrying humans who evacuated from Earth over seven centuries earlier. Suddenly, a tiny spark of electricity passes between them. WALL-E spins and pirouettes, giggling, flying around the spacecraft. EVE matches him move for move, in perfect unison. As both engage in synchronized flying around the ship, weaving in and out of rocket flames, harmoniously twirling double helixes, the gentle music in the soundtrack supports their dance-like movements by projecting a well-defined triple meter and tonality (D♭-Lydian). Humans aboard the Axiom watch in amazement as WALL-E and EVE spiral gracefully around one another. However, having lived for such a long time onboard the spacecraft, the humans no longer recognize or are physically capable of the earthbound activities their ancestors engaged in. Engrossed in Earth research, the captain asks the computer, “Define dancing”, to which it responds, “A series of movements, involving two partners, where speed and rhythm match harmoniously with music.” While the dialogue defines ‘dancing’, the audiovisual spectacle invites us to experience ‘dancing’. The music is thus integral in constructing this embodied experience—although underscoring zero-gravity visuals, the music affords us, and the passengers aboard the Axiom, an opportunity to feel grounded and engage in a virtual dance along with WALL-E and EVE.28 Through the music, we recognize what members of this futuristic society have lost and what they must now learn from the machines. Coda

Film music invites bodily engagement. As listeners, we subliminally react to the music’s properties and structure, attuning our perceptual experiences to its affordances. In this chapter, we explored the affordances of metrical and tonal frameworks, both in isolation and in combination, elucidating how these facilitate our embodiment of a film character’s physical actions and psychological states. In constructing a more ambitious argument, the chapter contemplates how the music’s affordances present an additional associative layer of meaning tied to the perceiver’s cultural practices and values.

Affordances

81

In Batman, characters lack superpowers. They obsessively seek, but ultimately fail, to transcend beyond their human boundaries. Batman resorts to material objects to extend his bodily capacities—armor to protect himself, gadgets to swiftly navigate his way through the gothic cityscape, a searchlight to project his image onto the smoggy skies and assert his presence—yet he is painfully aware of the limits of his human affordances. The music in the film emerges as one additional gadget—this time designed by the filmmakers to reach beyond the film’s storyworld—one that exists within an analogous mythical tension: the music subliminally forces us to resonate with the film’s characters in their attempts to reconstruct themselves yet remains trapped and constrained by their (and our) human affordances. Notes 1. Gibson (1979) argues that our perception of the environment is not based on a passive observation of its features, but on the active detection of the opportunities it affords for action. Through the notion of affordances, he acknowledges a continuum between perception and action, and a reciprocal relationship between an organism and its environment. Affordances therefore rest on the interplay between the environment’s structure, our capacities, and our intentions. A Gibsonian account of perception contends that we ‘resonate’ with our environments, recognizing ecologically relevant information; such an account shifts the focus from the properties of objects to the relationship between the observer and the environment. In turn, a broad understanding of the notion of ‘environment’ ensures that we recognize affordances as extending beyond the physical or material and onto the social and cultural realms. Music, for instance, when conceived of in terms of organism-environment interaction, offers ecologically relevant information, which we pick up from “the (sonic) environment and which affords perceptual significance” (Reybrouck, 2012, p. 394). For an in-depth discussion of the notion of affordances, see Appendix III in this volume. 2. Entraining to music is a complex phenomenon based on an interplay between the music’s hierarchical levels of patterning—rhythmic gestures, beats, (hyper)metrical units—and our bodily capacities. The beat is generally the most accessible (entrainable) temporal level, generally recognized by tapping to the music’s pulse, or ‘tactus’. Our perception of meter results from grouping beats into larger periodic temporal units; two-beats (duple meter) and three-beats (triple meter) groupings are the most typical. Because some listeners may hear slowly paced periodicities (e.g., at 30 BPM), while others may construct different perceptual impressions by entraining to nested periodicities, either twice as fast (at 60 BPM) or even four times as fast (at 120 BPM), within the context of this chapter I select periodicities that approximate 60 BPM. Furthermore, I presume that, like me, readers do not have access to the scores or other music-related information corresponding to the film examples, and therefore must rely upon their (innate and learned) capacities for entrainment. 3. Karl Bücher’s ethnomusicological investigations on work and rhythm point us to the origin of folk songs as the acoustic backdrop for the temporal structuring of physical labor. Meter in music that accompanies physical labor “necessarily results from our inner bodily constitution and from the technical preconditions of the work” (quoted in Meyer-Kalkus, 2007, p. 172). 4. In 1966, Pete and Toshi Seeger, alongside folklorist Bruce Jackson, produced a documentary about enslaved African Americans. They observed that songs would somewhat alleviate the strenuous physical effort and prevent individual workers from being singled out as performing more slowly than the rest. The documentary captures a real-life instance of collective entrainment to song, unfiltered by the aesthetics of a film director or composer. 5. Because quadruple meter contains a partially accented third beat ([1—2—3—4]–[1—2—3—4] . . .), we often perceive it as a variant of duple meter, therefore carrying nearly identical extramusical associations. 6. Duple meter is pervasive in filmic depictions of work and locomotion and in Western music writ large. Huron’s (2006) statistical studies of many musical genres illustrate that “duple and quadruple meters occur twice as often as triple and irregular meters (66% vs. 34%). In addition, simple

82 Affordances

7.

8. 9.

10. 11.

12.

13.

14.

meters are roughly six times more common than compound meters. In other words, . . . there exists a preference for binary beat groupings and a marked preference for binary beat subdivisions” (p. 195). He then builds upon the neurophysiological studies using EEG by Brochard et al.’s (2003) and on Jones and Boltz’s (1989) theory of rhythmic attention, to conclude that “there is some innate disposition toward binary temporal grouping” (p. 196). Theories of embodied cognition suggest that tempo perception originates in human locomotion (walking or running), where “the experience of rhythm is mediated by two complementary representations: a sensory representation of the motional-rhythmical properties of an external source, on the one hand, and a motor representation of the musculoskeletal system, on the other” (McAngus Todd et al., 1999, p. 26). In some unique cases, the music’s meter is employed as a form of word painting rather than to evoke a visceral response. A case in point is the film Stardust, where the music for the character Septimus incorporates a septuple meter ([1—2—3—4—5—6—7]–[1—2—3—4—5—6—7] . . .). Much research suggests that our tendency to entrainment is an evolutionary adaptation of our sensory systems. For instance, Phillips-Silver et al. (2010) note that entrainment builds upon “preexisting adaptations that allow organisms to perceive stimuli as rhythmic, to produce periodic stimuli, and to integrate the two using sensory feedback” (p. 3). Because entrainment allows for “successful predictions [that] can enhance the speed of perceptual organization” (Grahn & Rowe, 2009, p. 7547), our successful entrainment to cyclic patterns in our environments offers an evolutionary advantage. In turn, musical entrainment emerges as an extension of our broader tendency and capacity for entrainment. London (2012), for instance, speaks of entrainment in music as “a form of anticipatory behavior” (p. 25). Similarly, Large (2002) speculates that musical entrainment stems from “temporal expectancies that adapt in response to temporally fluctuating input” (p. 1). Therefore, in the context of adaptive cognition, temporal elements that undermine our capacity for entrainment, such as complex asymmetric meters, will elicit intensely unsettling physical and psychological responses. London (2012) draws a connection between temporal entrainment and the expressive power and social relevance of music, reminding us that we “learn to attune ourselves to the particular rhythms of our native musical culture[s]” (p. 64). Exploring entrainment as a “rhythmic dimension of human sociality”, for example, McNeill (1997) points out that synchronized marching in combat serves a ‘bonding’ rather than a mere ‘coordinating’ function. Furthermore, he suggests that entrainment as a means for social bonding and identity formation may extend beyond marching, and remarks that for the Aztecs, the practice of dancing ensured that “warriors asserted their corporate identity and prepared for battle” (p. 103). Keyfitz (1996), in turn, suggests there is a lasting psycho-social impact to communal entrainment, speculating that whether “ambushing a tiger in the hunt [or] attacking an enemy stockade, the group that has drilled or danced together is more likely to come out successful through the mutual trust among its members developed as bonding” (p. 408). Similarly, in a scene from Farewell My Lovely [00:44:10] the film’s femme fatale and a private investigator speak alone. The music enters as she says, “Why don’t you come over here and sit beside me”, to which he responds, “I’ve been thinking about that for some time, ever since you first crossed your legs.” In this scene, the visuals suggest analogous dynamics to Punch Drunk Love between the characters—by their position opposite to each other and by the color of their clothes—while the music in triple meter underscores a similar dance of seduction and anticipates their becoming intimate. Even when asserting that associations triggered by physiological responses are ‘learned’, the learning process is a biological (not a disembodied) phenomenon based on motor and neurological mechanisms that stem (in this case) from temporal patterns of rhythmically coordinated experiences. The visuals support and contribute to this interpretation. Initially framing Jenny and a luxury car on opposite sides emphasizes Jenny’s distance from the lifestyle David offers. Subsequently, as Jenny no longer resists the idea of engaging in unethical behavior, David swings her over to the side of the frame where the luxury car is located, suggesting a transformation that aligns with a different set of values. As David and Jenny continue dancing, the sudden appearance of two luxury cars aligned on the visual frame (and matching the colors of David’s and Jenny’s attire) further supports this highly suggestive visual metaphor.

Affordances

83

15. When conducted on Earth, this research simulates weightlessness conditions; for a review see Oman (2003). 16. In space, for instance, the conflict between proprioceptive information and visual or tactile cues creates a “disparity between sensory input from various sources [that] may result in acute disorientation” (Tischler & Morey-Holton, 1992, p. 1345). For additional sensory conflict models see Oman (1998); Reason and Brand (1975). 17. For instance, the “sensory conflict produced by the altered gravireceptor signals produces symptoms of motion sickness” (Oman, 2003, p. 202), mainly when “vestibular stimulation is in conflict with all other sensory information” (Hecht et al., 2001, p. 116). 18. The perception of a tonal center is one of the most complex phenomena studied in music cognition and the subject of extensive experimental research. Browne’s (1981) “rare interval” hypothesis holds that a scale’s unique intervallic profile contributes to defining a pitch collection’s qualia in terms of its gravitational center. Symmetrical collections are characterized by cyclical intervallic patterns, high frequency of a small number of intervals, and sub-segments that map onto themselves verbatim under various degrees of transposition or inversion. The absence of statistically rare intervals and the absence of unequal distribution of intervallic structures in symmetrical collections is foundational to our perception of their qualia. Nevertheless, various compositional strategies—the archetypical gestures and style-bound conventions embedded in the musical texture—play an equally significant role in eliciting various degrees of tonal centricity; as Butler (1989) notes, the “listeners’ judgments of tonal center are strongly influenced when rare intervals are arranged differently across time” (p. 234). 19. In contrast, the ‘rare’ tritone and diminished triad in a major scale contribute to orienting us toward the scale’s gravitational center (or tonic) by eliciting a set of tendencies in relation to the tonic as the most stable point. Therefore, the absence of rare intervals or rare triads in symmetrical collections does not afford us a landmark, a pointer, to orient us toward the collection’s gravitational center. 20. The visuals of space expeditions often seek to elicit feelings akin to weightlessness by simulating zero-gravity environments, triggering in us a cognitive dissonance between the optical perception of floating and the proprioceptive sensation of gravity. For instance, in exploring such visual techniques, D’Aloia (2012) argues that the spectator “loses (at least temporarily) the sensation of being grounded to a surface of support . . . in the immersive darkness of the movie theatre, the spectator senses an incongruity between what is seen and what is felt” (p. 232). In turn, the music is a powerful device, one that operates in concert with other cinematographic techniques. In fact, research shows that our motor behavior responds more strongly to aural than visual stimuli (see Patel et al., 2005; Repp & Penel, 2004). In scenes depicting space expeditions, for example, music that draws on symmetrical pitch collections may engender this inter-sensory conflict, thereby further enhancing the cognitive dissonance. 21. A tritone, for instance, embeds an unresolved sensory dissonance that causes a “physiological interference along the basilar membrane of the cochlea . . . [rendering] the hearing organ less able to discern the various spectral components present in the environment” (Huron, 2006, pp. 324–325). 22. Murphy (2006) speaks to the use of the tritone, and specifically the major tritone progression, in recent Hollywood films featuring scenes set in outer space. 23. In a Shepard tone, the superposition and loudness fluctuations of sine waves an octave apart create the auditory illusion of a continually ascending or descending pitch. 24. In Cloudy with a Chance of Meatballs, a descending sequence of dominant-seventh harmonies underscores the protagonist’s wandering in the air at the mercy of a strong tornado. [00:50:20] Each dominant-seventh harmony is devoid of its resolution to the tonic, and instead, each leads to another dominant-seventh harmony a half-step below it. This descending chromatic sequence systematically thwarts the listener’s tonal expectations, not affording the listener a tonal center. Only as the protagonist reaches the ground, the music resolves to a stable sonority that functions as a tonic, denoting that the protagonist reached firm ground. Scenes depicting falling characters or objects typically blend the LINEARITY image schema (see Chapter 3) with a musical dissolution of gravitational forces. 25. The different ratios of intervallic constructs embedded within pitch collections prompt us to construct mental representations (i.e., schemas) particular to each pitch collection—the distances

84

Affordances

between adjacent tones in a collection gives “rise to the emergence of the very different qualia” (Shepard, 2009). 26. Film directors often draw on scientific research and new technologies to arrive at cinematographic artifices that elicit in us specific responses. For instance, in 2001: A Space Odyssey, Kubrick pioneered a visual effect to simulate a zero-gravity environment; to achieve this, he designed a circular set that rotated vertically and instructed the actors to move at the rotating speed so as to remain at the bottom of the set. 27. However, as the film unfolds, this normalcy fades away, and the music becomes dissonant and ametric, unfamiliar and alien. 28. Oden (2023) explores how triple and compound meters generate a feeling of weightlessness, including in this scene from WALL-E.

6 MEMORY & AUDITORY PERCEPTION

In Jaws, the music lurks from beneath our conscious attention and conditions us to respond automatically, reflexively, instinctively. The film opens with a dozen young men and women gathered around a beach bonfire, serenely trading songs. Chrissie and Tom break away from the circle and head for a swim. Chrissie runs up a dune and pauses to look at the quiet ocean, the sun setting on the horizon. She tosses her clothes on the sand and placidly draws herself deeper and deeper into the soothing, silent waters. Tom remains on the shore. The music sneaks into the soundtrack, with a two-note figure emerging from the depths of the orchestra. A ripple in the water grabs Chrissie’s attention as she feels gently pushed up and pulled down. No, it is not Tom playing around; he is still on the beach. Frightened, she swims toward the shore, but the ripple moves with her. Now, with a tenacious drive, the twonote musical figure intensifies, sucking Chrissie (and us) down to her horrific fate. A leitmotif has emerged. Leitmotifs are more than memory traces. They are to us what the sound of a bell is to Pavlov’s dog: ecologically relevant acoustic cues that rely on evolutionary mechanisms to condition us to respond in specific ways.1 During a film, leitmotifs manifest themselves through the concurrent and consistent appearance of particular musical figures and their extramusical counterpart; once established, they effectively bring to mind their extramusical counterparts and trigger a visceral response. In this chapter, we first delve into the various rationales behind leitmotif construction. Then, we attend to the cognitive mechanisms that allow us to identify these figures, solidify them in our minds, and associate them with their extramusical counterparts. Last, we reflect on the entailments of leitmotifs from the vantage point of evolutionary psychology. Leitmotif Construction

Leitmotifs straddle both the embodied and semiotic worlds, eliciting physiological responses while also serving as musical signs. This dual nature of leitmotifs, which imbues them with unmatched rhetorical and expressive power, manifests itself during the compositional

DOI: 10.4324/9780429504457-7

86 Memory & Auditory Perception

process. To design leitmotifs with robust connections to their extramusical counterparts, composers construct musical figures that elicit in us hardwired responses, that draw on the music’s affordances (and the embodied associations these bring about), that acoustically resemble elements in the storyworld, or that evoke their extramusical counterpart via learned socio-cultural associations. Hardwired Responses

Composers design leitmotifs that elicit hardwired responses by establishing analogies between the visual and aural domains. In these cases, the extramusical counterpart’s physiognomy (such as size, shape, density, reflectance, texture) helps determine the musical characteristics (such as loudness, pitch range, timbre, contour). In tracing our affective responses to these musical characteristics, we recognize that our evolutionary history has hardwired us to respond in specific ways. When foregrounded within leitmotifs, these musical characteristics elicit visceral responses analogous to those elicited by the leitmotif’s extramusical counterpart. Leitmotifs can therefore indicate their extramusical counterpart’s temperament and physical appearance—small and friendly characters or objects would feature a small ensemble sound playing softly, whereas large and threatening characters or objects would feature large symphonic forces playing loudly. In Maleficent’s opening scene, we learn that in a magical kingdom lived a wonderful creature, a fairy girl, whose name was Maleficent. A young girl with small wings rests happily on a tree branch, innocently playing with dolls and gently wrapping her tiny hands around a broken tree branch to heal it. At this tender moment, the soundtrack introduces young Maleficent’s leitmotif—a gentle melody on resonant instruments performing in their

FIGURE 6.1A

‘Young Maleficent’ leitmotif. (Music by James Newton Howard.)

FIGURE 6.1B

Maleficent. Young Maleficent innocently plays. [00:01:00]

Memory & Auditory Perception

FIGURE 6.2A

‘Fierce Maleficent’ leitmotif. (Music by James Newton Howard.)

FIGURE 6.2B

Maleficent. Maleficent turns to rage and revenge. [00:25:10]

87

high registers. Later in the film, a much older Maleficent learns of a betrayal that hardens her heart. She screams with rage and marches toward the center of the kingdom. White fire erupts from her, scorching everything in her path. Maleficent now has turned into a fierce creature bent on revenge. During this threatening moment in the film, the soundtrack introduces a different leitmotif—densely stacked chords in the low register performed by a large ensemble that includes nearly the entire orchestra—to represent the now-unkind, fierce Maleficent. In both instances, the leitmotifs’ musical characteristics directly tap into our evolutionary, hardwired responses. Music’s Affordances

Film composers often rely on the music’s affordances (and the embodied responses these elicit) to establish a strong link between a musical figure and its extramusical counterpart. Such affordances transpire through a leitmotif’s structural parameters—as we embody these structural parameters, we subliminally engage in a broad array of interactions with the music, interactions that directly relate to a leitmotif’s extramusical counterpart.

88 Memory & Auditory Perception

FIGURE 6.3A

‘Vertigo’ leitmotif. (Music by Bernard Herrmann.)

FIGURE 6.3B

Vertigo. Scottie follows Madeleine up the bell tower. [01:16:00]

Vertigo illustrates a masterful interplay between various cinematic devices and the audience’s reactions to those devices. Scottie, who suffers from an intense fear of heights, attempts to prevent Madeleine from committing suicide. Steps from a tall church, they exchange tender words. Madeleine, however, seems distressed; she runs to the bell tower and quickly goes up the stairs. Scottie follows her but cannot keep up—his vertigo paralyzes him. At this moment in the soundtrack, and at similar moments throughout the film, the music signals the protagonist’s outbursts of vertigo with a distinctive leitmotif that affords us no firm tonal grounding—a chordal construct superimposing two unrelated triads (E♭m and D) and harp sweeps in two different modes (E♭-Aeolian and D-Ionian), forming a near-dodecaphonic collection—depicting (and eliciting in us) a feeling of bodily and psychological instability.2 In The Matrix, characters bend time and space. Trinity is on the run, with armed police officers and two agents following her. She emerges as a black blur in the dark cityscape, darting across rooftops as if she were part of the shadows themselves. The police chasing her seem heavy and lumbering in comparison, but two agents take the lead, almost matching her athleticism. With a sudden burst of speed, Trinity reaches the edge of a rooftop and defies gravity, leaping in the air, her eyes fixed on the distant building. Time seems to stretch out endlessly. As she hurtles through space, the wind whips past her face, the sounds of the city fade away, and the ‘Trinity leap’ leitmotif soars in the soundtrack:

Memory & Auditory Perception

FIGURE 6.4A

‘Trinity leap’ leitmotif. (Music by Don Davis.)

FIGURE 6.4B

The Matrix. Trinity on the run. [00:04:30]

89

a set of oscillating triads that disrupt the firmly grounded musical gravitational forces and an (a)metric profile that suspends musical temporality. Then, with a soft thud, she lands on the opposite rooftop, her form perfect and controlled. The agents, left behind on the first rooftop, stare in disbelief at Trinity’s audacious feat. Throughout the film, the ‘Trinity leaps’ leitmotif capitalizes on our tendencies to attune to the music’s affordances with a figure that evokes a suspended embodied response analogous to Trinity’s gravity-defying moves. Acoustic Resemblance

Via musical onomatopoeia—imitating the sound of an extramusical element—composers construct leitmotifs that feature a close acoustic resemblance between the musical and extramusical elements. By mimicking a character’s sounds and thereby drawing on our pre-existing auditory associations, leitmotifs begin to extend into the terrain of musical semiotics.3 Return to Oz presents numerous leitmotifs, most of them resting on the associative powers of onomatopoeia. The leitmotif for Billina, a chicken who travels with Dorothy, captures the nasal timbre of a chicken’s cackle as well as the chicken-esque intervallic patterns and articulation; while the two oboes and a bassoon help construct Billina’s timbric characteristics, the angular-shaped (and somewhat disjointed) musical figures provide the contour and articulation that contribute to onomatopoeically capturing her cackle. The leitmotif for another character, Tik-Tok, a metallic mechanical character, draws primarily on timbre; it features the metallic tone of an all-brass ensemble formed by one cornet, two baritone horns, and one tuba.

90 Memory & Auditory Perception

FIGURE 6.5A

‘Billina’ leitmotif. (Music by David Shire.)

FIGURE 6.5B

Return to Oz. Billina appears. [00:22:50]

FIGURE 6.6A

‘Tik-Tok’ leitmotif. (Music by David Shire.)

FIGURE 6.6B

Return to Oz. Tik-Tok comes alive. [00:33:00]

Memory & Auditory Perception

FIGURE 6.7

91

The Terminator. T-1000 emerges unscathed. [00:32:00]

When an extramusical counterpart resists onomatopoeic representation—because the sounds it makes are unknown or absent—composers devise ingenious musical solutions. In The Terminator, a cyborg disguised as human travels from 2029 to 1984 to kill a young woman whose future son is the key to humanity’s salvation. Constructing the leitmotif for the main character, the Terminator, presented the composer with an opportunity to envision the sonic signature of a cyborg. The musical figure, an electrostatic rumble in the low register, becomes a fantastical onomatopoeic rendition of a cyborg’s reverberating mechanical parts. Musical Archetypes

Musical ‘topics’ function as archetypes, providing composers a scaffolding framework to elicit, through their music, specific social and cultural associations.4 When weaving musical topics through the fabric of leitmotifs, composers are keen to foreground a topic’s characteristic ‘marker(s)’. Although any musical parameter may potentially amount to a defining topical marker, pitch collections (in the form of modes or scales) are particularly effective and versatile. Kung Fu Panda illustrates the use of a pitch collection, while also including culturally specific instrumentation as an equally effective topical marker. This animated film unfolds in the Valley of Peace, a land in ancient China inhabited by anthropomorphic animals. The ‘Hero’ leitmotif for the main character, Po, who secretly dreams of becoming a Kung Fu legend, draws on a pentatonic scale [D, F, G, A, C] and uses strings and traditional Chinese instrument, to signify his origins.5 This leitmotif emerges for the first time as Po dreams of saving his village from the evil Tai Lung.

FIGURE 6.8A

‘Heroic Po’ leitmotif. (Music by John Powell).

FIGURE 6.8B

Kung Fu Panda. Po’s dream of greatness. [00:01:15]

92 Memory & Auditory Perception

FIGURE 6.9

‘Darth Vader’ leitmotif. (Music by John Williams.)

Often, numerous markers combine to define a topic. In the Star Wars films, Darth Vader’s leitmotif’s minor mode, dotted rhythm, and slow tempo evoke the ‘funeral march’ topic, along with its socio-cultural associations.6 Here, however, multiple additional mechanisms play a role in strengthening the link between the leitmotif and its extramusical counterpart: its low register and relatively high loudness fill the entire aural space, instilling in us the fear that stems from encountering an overpowering large force; its duple meter engages our embodied affordances of marching, bringing about resonances of war and conflict;7 and, its oscillating chromatic-third relationships [C-minor, A♭-minor] suggest a complementary topic associated with the character’s sinister nature.8 Leitmotif (Re)Cognition

For a musical figure to become a leitmotif, listeners must engage in numerous cognitive mechanisms related to perception and memorization: we subliminally embody a musical figure via subvocalization, organize and decode it according to gestalt principles, solidify it in memory through subvocal rehearsal, and categorize it according to learned musical schemas.9 Such perception and memorization mechanisms are vital in setting the space for a subsequent associative mechanism, where we link a ‘proto’-leitmotif to an extramusical element. Ultimately, once established, a leitmotif’s appearance within the fabric of a film’s soundtrack will elicit the presence of its associated extramusical element, thereby reinforcing a leitmotif’s semiotic nature as a musical sign. In this section, we draw exclusively on Jaws’ famous leitmotif to illustrate how this multipronged mechanism subliminally leads us from hearing a two-note growling musical figure to sensing the presence (virtual or realized) of a menacing shark. Perception

Jaws’ opening title sequence submerges us within an underwater point-of-view that suggests the perspective of an animal lurking in the ocean’s depths. In the soundtrack, the unnerving silence is gradually filled with a two-note figure in the low register. Our brains and bodies actively (yet subliminally) respond to the low rumblings in the music: we subvocalize the ascending semitone musical figure, entrain to the regularity of its underlying pulse, and increase our heart rate and pupil dilation in response to its unpredictable accents. We also engage gestalt principles (italicized here) to decode this acoustic stimulus. The initial temporal and registral proximity of two notes (E – F) and the silences between iterations create the impression of a single melodic figure. A gradual unfolding of this figure via its obsessive repetition suggests an embryonic musical process—the similarity of all repetitions,

Memory & Auditory Perception

FIGURE 6.10

Jaws. Opening sequence.

FIGURE 6.11

‘Jaws’ leitmotif, measures 1–10. (Music by John Williams.)

93

emerging with an ever-increasing frequency and within a clear metrical structure, prompts us to perceive a lengthier, fully formed musical theme. (See Figure 6.11.) Its growth does not stop. This two-note figure further expands harmonically into two semitone-oscillating dense harmonic constructs that, with their analogous contours and simultaneous onsets, engage the common fate principle, prompting us to group the various layers of this sonic construct.10 (See Figure 6.12.) While this two-note figure continues growing, the music presents a secondary melodic figure—it is precisely the proximity, similarity, and common fate gestalt principles that allow us to identify this as a ‘secondary’ figure, one that complements (yet is dissociated from) the primary one. Toward the end of the opening titles sequence, the initial two-note figure gains further strength, adding octave doublings, capturing the entire register.11 By the end of the main title sequence, listeners have subliminally processed the two-note musical figure to the extent

FIGURE 6.12

‘Jaws’ leitmotif, measures 16–18. (Music by John Williams.)

94 Memory & Auditory Perception

FIGURE 6.13

‘Jaws’ leitmotif, measures 20–22. (Music by John Williams.)

that it has become a proto-leitmotif—a sonic object with the potential to become a fullfledged leitmotif that will hook us to the story. (See Figure 6.13.) To make a musical idea recognizable (and potentially more memorable), composers embed ‘salient’ features in one (or more) of its parameters—for instance, an unfamiliar timbre, an unusual rhythmic gesture, or an uncommon contour.12 In doing so, composers intuitively draw on our evolutionary history, on our hardwired tendencies to respond to changes in an otherwise regular environment.13 In Jaws, the unusual combination of contrabassoon and double-bass performing a motivic idea in the instruments’ lowest registers (rather than an accompaniment figure or a bass-line) cuts through the listeners’ perceptual habituation.14 Memorization

Leitmotifs are fluid, idealized mental constructs, morphing with every new instance, always containing within themselves the potential for multiple expressions.15 Our ability to solidify a leitmotif in short-term memory is contingent upon the interplay between learned schemas (entailing a top-down cognitive mechanism) and its repetition (both, embodied repetitions in the form of subvocal rehearsal, and actualized repetitions in the form of iterations in the soundtrack).16 Learned schemas serve to offload cognitive processing and provide a scaffolding framework for perceiving information.17 These mental constructs are vast and varied, spanning from phonological to syntactical musical attributes, including tuning systems, pitch collections, timbres, metrical patterns, phrase structures, and even stylistic or performative traits. When subliminally parsing new music stimuli through learned schemas, we further ‘chunk’ information and identify schematic correspondences.18 Music that draws on encultured listeners’ learned schemas—e.g., using a familiar tuning system or drawing on archetypical gestures of a recognizable style—is more prone to stick in the memory.19 For this film, anecdotal evidence suggests that the primary figure, the two-note chromatic ascent, firmly remains in the collective consciousness as ‘the Jaws theme’, while the secondary figure, the arpeggiated gestures in the horns, is seldom if ever recognized in connection to the film. This difference results, in part, from schemas at work: tonally, the primary figure easily relates to learned melodic and tonal schemas, whereas the secondary figure does not; and temporally, the primary figure projects a clear metrical pattern, becoming a driving ostinato that marks the pulse, thus nicely fitting within the temporal scaffolds of learned schemas, whereas the secondary figure is not bound by the established

Memory & Auditory Perception

95

quadruple meter, freely introducing irregular durations (triplets, quintuplets, septuplets) that further divorce it from the scaffolds of common metrical schemas.20 Therefore, the primary figure easily sticks in our memories, while the secondary one quickly fades. Mere exposure to an acoustic stimulus does not ensure its memorization. Initial ‘echoic memory’ becomes extended through the ‘phonological loop’, a covert and subliminal repetition mechanism that engages subvocal articulation and rehearsal and that helps us memorize auditory stimuli. In the context of Jaws, although the initial two-note gesture ceases to sound, we continue to subvocalize it, which subliminally contributes to its memorization. The importance of the phonological loop notwithstanding, forming a robust memory trace may only result through actualized repetitions in the soundtrack. Just as in all types of cognition, the number of repetitions of a stimulus and the potentially interfering informational ‘noise’ in between repetitions significantly affect the perception and memorization processes.21 Certain auditory conditions set the space for effective perception and memorization—to promote distinct memory traces, musical figures must appear isolated from distracting auditory signals (dialogue, sound effects, or even busy accompaniment textures) and must repeat within short timespans.22 In Jaws, these two conditions are met. At the onset of the main title sequence, the two-note musical figure emerges as the only sound element in the film, unhindered by interfering dialogue or sound effects. Less than three minutes into the film [00:02:30], a new music cue repeats identical materials, only competing with sporadic cries for help from the victim. Shortly after [00:15:30], yet another cue presents identical musical materials, but now completely isolated, not hampered by any other soundtrack layer. Association

Through the phases described up to this point, we have perceived acoustic information, recognized units within it, and stored these in short-term memory. However, for these musical units to function akin to stimuli in classical conditioning, they must be associated with an extramusical counterpart in the storyworld: during a film, a musical figure consistently and systematically appears coupled with an extramusical element in the storyworld (an object, a character, a situation), and after repeated exposure, the appearance of this musical figure within the film’s soundtrack will suggest the presence of the associated counterpart in the storyworld.23 By virtue of their association with a secondary stimulus, leitmotifs take on a semantic nature, therefore residing along a continuum between implicit and explicit memory.24 In Jaws, the concurrent presence of the characteristic two-note ascending chromatic figure and the idea of a shark surface in various ways. During the main title sequence, the visuals subliminally imply the presence of the extramusical counterpart, assuming a visual perspective from the ocean’s depths. Shortly after, in the second cue, Chrissie’s bizarre movements suggest the extramusical counterpart’s presence as she is violently tossed from side to side and ultimately pulled under the water. By the time we hear the third cue, the dialogue has revealed the presence of a shark in the waters; while there has as yet been no visual confirmation of the existence of a shark, we witness the outcome—blood suddenly surges in the water where children are playing and swimming. It is not until the fourth cue that the music finally emerges coupled with images of the shark, this time circling a character submerged in a seemingly flimsy cage. By this point in the film, the leitmotif is well-established. Now,

96 Memory & Auditory Perception

the expressive and signifying powers of the leitmotif emerge powerfully—by drawing on our hardwired tendency to classical conditioning, the soundtrack introduces subsequent iterations of this leitmotif to signal the dangerous shark’s imminent or latent presence. Evolutionary Function of Leitmotifs

Auditory perception, memory, and associative mechanisms support vital evolutionary and ecological functions. Associating an auditory signal with an extra-auditory event results in adaptive behavior—e.g., a voice calling out our name grabs our attention, a doorbell alerts us to someone at the door, and the sound of a car approaching prompts us to move to the sidewalk.25 Such vital functions of auditory perception and memory remain engaged within all human milieus, even within culturally constructed spaces that seem distant from natural environments, such as a movie theater. Embedded within the musical fabric of a film’s soundtrack, leitmotifs operate just like other environmental sounds, triggering psychophysiological responses and adaptive behavior.26 Our ecologically resonant minds and bodies treat leitmotifs like any other environmental sounds: once we associate a musical figure with its extramusical counterpart, either stimulus in isolation will trigger one and the same response.27 Jaws’ leitmotif capitalizes on the evolutionary and ecological roles of auditory perception and cognition. Upon the film’s release, Jaws’ leitmotif became instantly associated with the shark in the film, as well as with our strong emotional responses to shark attacks and with a conglomerate of contextual cues from outside the film, including dangerous situations and even the shallow waters of the beach.28 This is evidence that leitmotifs exploit our ecological and evolutionary tendencies in the theater and outside of it, creating long-lasting ecological and behavioral responses akin to classical conditioning, permeating real-life settings and becoming part of our cultural milieus.29 Coda

Film music leitmotifs signal the presence of extramusical elements—characters, places, events, or other narrative elements of a film’s storyworld. This chapter begins to reveal the multi-level complexity of leitmotifs: while on the surface they appear as intraopus-established symbols, at a deeper level, they function as ecologically relevant acoustic cues that rely on the listener’s evolutionary, hardwired mechanisms. Jaws navigates through a space of cultural significance while sinking its teeth into our psyche. While the film’s surface is flooded with symbolism that grows more prophetic over time, its lurking leitmotif operates beneath our conscious attention, ultimately altering our behavior and relationship with nature, leaving a lasting, indelible yet unseen scar. Notes 1. Leitmotifs have enjoyed much attention from film musicology; in fact, identifying and labeling leitmotifs has long been a staple of film music scholarship. Although some recent musicological studies on leitmotifs are informed by cognitive psychology, these do not engage in the empirical investigation and thus remain firmly rooted in the humanities. These include the exploration of prototype formation (Bribitzer-Stull, 2015; Zbikowski, 2002), memory (Biancorosso, 2013; Cohen, 2014), evolutionary perspective (Biancorosso, 2010), and cue abstraction and imprint formation

Memory & Auditory Perception

2. 3. 4. 5. 6.

7.

8. 9.

10. 11. 12.

13. 14.

97

(Cambouropoulos, 2001; Reybrouck, 2010; Wiggins, 2010). Additionally, scholars apply methodologies from cognitive sciences to understand leitmotifs from behavioral and psychological perspectives, with studies focusing on perception (Baker & Müllensiefen, 2017), memory (Boltz et al., 1991, 2009), cognitive processing (Boltz, 2001, 2004; Hacohen & Wagner, 1997; Nosal et al., 2016; Tan et al., 2010; Töpper & Schwan, 2008), cue abstraction and imprint formation (Deliège, 1992, 2001; Deliège & Mélen, 1997), salience (Deliège et al., 1996), and physiological responses (Cantor, 2004). For in-depth discussions of memory and auditory perception, see Appendices IV and V in this volume. Murphy (2022) observes correspondences between the music and the visual artwork presented during the main title sequence. Since the relationship between signifier (the musical sound) and signified (the actual sound) is based upon similarity, musical onomatopoeia amounts to a semiotic function—primarily an iconic sign within the context of Peirce’s trivium. Chapter 7 offers an in-depth treatment of film musical ‘topics’. In the animated film Lady and the Tramp, the Siamese cats’ leitmotif outlines a pentatonic scale to signify their Asian ancestry. Similarly, in Dr. No, the pentatonic scale is used for the protagonist’s leitmotif to connote his Chinese descent. Most funeral marches in the common practice repertoire feature analogous parameters. See for instance, Chopin’s Bb-minor Piano Sonata, Beethoven’s Symphony #3, Grieg’s Funeral March, Berlioz’s Marche Funèbre Pour la Dernière Scène, Mahler’s Symphony #5, Beethoven’s Piano Sonata #12, and Alkan’s Funeral March on the Death of a Par. By drawing on topical signification, some leitmotifs (tangentially) embed traces of embodied affordances to suggest character traits. For instance, comparing the meter of leitmotifs representing two villains, Darth Vader from Star Wars and Joker from Batman, sheds light on the musical characteristics that reflect their individualities—whereas the duple meter in Darth Vader’s leitmotif represents his martial and warlike nature, the triple meter in Joker’s leitmotif represents his odd, circus-like appearance. Lehman (2018) defines this specific chromatic-third relationship as evoking “a feeling of harmonic unnaturalness . . . the affective ‘dark side’ ” (p. 101). Leitmotif cognition engages bottom-up and top-down mechanisms. These are functionally distinct: while bottom-up mechanisms are geared toward constructing perceptual entities, top-down mechanisms are geared toward selecting information by activating learned schemas. Reybrouck (2010) echoes this duality in perception and distinguishes between ‘extracting’ salient features from the musical surface and ‘abstracting’ the musical surface; while the former relates to the sensory experience of music according to gestalt principles, the latter relates to the mechanisms of cue-abstraction and imprint-formation necessary to the construction of a learned schema. For some listeners, this musical gesture is reminiscent of Stravinsky’s The Rite of Spring chord. Here, the primary and secondary musical figures blend: the harmonic construct draws on a pitch collection associated with the secondary theme (E, G, B♭, D♭, E♭), yet its voicing foregrounds interval-class 1 between its outer notes, the primary interval of the primary two-note figure. Deliège (1992) attends to “head structures” (the initial portion of a theme) and passages with “accents” (musically marked, salient events in terms of loudness, range, contour, temporal placement, duration, or any musical parameter), and suggests that both become imprinted in memory to a greater degree than the remaining portions of a melody. There is, therefore, a delicate balance between abiding by common schemata and attaining saliency; to construct memorable and recognizable musical figures, composers must do both: draw on common musical schemas and include salient features. The saliency of a proto-leitmotif’s initial presentation is vital to ensure its recognition, especially because subsequent iterations may be obscured in the soundtrack, buried under dense accompaniment textures—successful recognition of subsequent iterations is contingent upon our forming a strong imprint of the proto-leitmotif. Much of Dowling’s influential research program attends to the interaction of contour (bottom-up mechanisms akin to gestalt) and scale/interval information (topdown mechanisms grounded in schemas) in the memorization of melodies; in several studies, he suggests that when exposed to short, atonal, or novel melodies, we are more prone to retain their contour information, whereas when exposed to longer, tonal, or familiar melodies, we are more prone to retain a melody’s interval information (see Dowling, 1978; Dowling & Bartlett,1981; Dowling & Fujitani, 1971). More recent neuropsychological studies also suggest the primacy of

98 Memory & Auditory Perception

15.

16.

17.

18.

19.

20.

21. 22. 23.

contour in (subliminally) discriminating melodies; for instance, using electroencephalography (EEG) to measure event-related potentials (ERPs), Schiavetto et al. (1999) found more significant (larger and earlier) neuronal responses to contour changes than interval changes in short melodies. While exploring a phenomenon related to music and memory, Deliège and Mélen (1997) argue that ‘imprints’ (mental constructs of melodies) are not “a set of fixed traces [but instead] a central yet moving and flexible tendency which either settles or readjusts depending on the particular presentation of the cues” (p. 403). Most research attends to the familiarity or lack of familiarity with melodies by categorizing these respectively as short-term or long-term memory traces. Although most leitmotifs fit within shortterm memory (as novel melodies introduced within the context of a film), some leitmotifs turn into long-term memory traces, particularly within franchises such as Star Wars or Lord of the Rings. Krumhansl (1991) mentions that listeners use “their knowledge of the interval patterns typical of the style to apprehend and retain larger and more musically meaningful patterns” (p. 295) and cites various studies confirming that diatonic scale structures play a significant role in memorization and recognition, particularly in transposed melodies. Deliège (2001) notes that, when constructing mental images of leitmotifs, our “memory simplifies and reduces the global information, developing statistical means . . . that retain the essential information about a collection of more or less similar presentations” (p. 372). Deliège and Mélen (1997) provide a detailed account of this process through the notions of cue abstraction and imprint formation: instead of storing individual exemplars, listeners attend to ‘cues’—salient, surface-level, marked, qualitative characteristics in the music—which are stored in working memory; subsequently, ‘imprints’ are drawn from “the traces deposited by the accumulation of varied repetitions of cues [and] are transformed into a sort of ‘résumé’ of the . . . basic structure” (p. 403). Hence, while cue abstraction is a mechanism that progressively processes the musical surface by tracing marked and salient differences, imprint formation is driven by a progressive (proto-statistical) recognition of similarities among cues that unfold diachronically throughout an entire film. However, music that abides too strongly by learned schemas may go unnoticed—most background music during a film remains ‘unheard’ (to borrow Grobman’s term) precisely because it is highly generic. Biancorosso (2013) speaks to this phenomenon and notes that generic music embeds musical figures “whose sheer commonality renders them forgettable and whose evocative purchase has been wholly exhausted” (p. 206). Dowling (1972) observes that atonal melodies “are typically difficult to deal with in recognition experiments because of their departure from the well-learned scale functions the listener hears in the rest of [their] musical experience” (p. 420), suggesting that tonal melodies may be more readily recognized than atonal ones because listeners rely on schematic tonal information. In addition, Jones (1993) attends to the temporal aspects of music and argues that memorization of melodies relies on “dynamic pattern structures” that emerge from combining changes in pitch space (such as contour or interval sizes) with changes in the temporal domain (such as durations). Byron (2008) echoes these thoughts and notes that “memory for melodic contour and memory for rhythmic structure interact in various ways . . . [suggesting] that both duration information and pitch information are used in the chunking of melodies that must occur as a result of short term memory processes which are limited in the amount of chunks that can be held at any one time” (p. 78). For instance, learning words in a new language is most efficient when hearing them repeatedly within short time spans and isolated from other confounding words or sounds. Main title sequences at the onset of a film are thus the ideal film segments to introduce leitmotifs because the music does not compete for space with sound effects or dialogue in the auditory landscape. In a study exploring cross-modal attention, Boltz (2004) argues that “music and film are jointly encoded as a unified entity” and that “attending to one dimension should result in the incidental learning of the other, and no performance decrement should occur if attending is divided between both dimensions at once” (p. 1195). In contrast, when discussing the use of leitmotifs in opera, some scholars suggest that the visuals hinder the memorization of leitmotifs; for instance, Baker and Müllensiefen (2017) note that “seeing and hearing the opera actually decreased an individual’s ability to identify leitmotives in the auditory signal and hence suggests that visual information can act as a distractor in terms of encoding leitmotivs” (p. 2).

Memory & Auditory Perception

99

24. Although most research on melody recall and discrimination examines the role of conscious attention, leitmotifs are embedded within the musical fabric, which often goes unnoticed and is seldom attended consciously. 25. Our ability to create and recognize patterns in music may have been a precursor to our ability to create and understand language, as both involve organizing and structuring sounds. Therefore, our tendencies to recognize and memorize leitmotifs may rely on a pre-linguistic evolutionary function. 26. Rötter (1994) identified marked physiological reactions, such as peaks in Galvanic skin response, that coincide with the occurrences of leitmotifs, suggesting a phenomenon akin to classical conditioning. Biancorosso (2010) speculates about Jaws’ leitmotif triggering such classical conditioning, noting that “by the time the audience (or those viewers already in the know) sees the subsequent attacks, the connection between the minor-second motive and the action becomes Pavlovian in its reflexivity” (p. 307). 27. Often, salient musical gestures that do not function as leitmotifs become so intricately associated with an element in the storyworld that they transcend the boundaries of their original film and find their way into other films as a quotation or parodic commentary—the shrieking string gestures from Psycho explored in Chapter 1 are a prime example. 28. Most scholars regard Jaws’ as the leitmotif par excellence because it uses “mood-congruent relationships . . . [to] reflect both a joint encoding and a unified memory representation of music and film information” (Boltz, 2004, p. 1196) and because it relies on “events associated with strong emotions . . . [which] are better remembered than emotionally more neutral events” (Guenther, 2002, p. 323). In fact, extra-musical elements such as the narrative, context, associations, and other sensory information all contribute to our (re)cognizing and solidifying melodies in our memory, particularly when there is structural, affective, or semantic congruency or alignment between the stimuli across modalities. Tulving’s (1983) ‘encoding specificity’ hypothesis purports that the presence of contextual similarity surrounding the initial encoding and subsequent retrieval of memories helps consolidate them as long-term memory traces. This suggests that reinforcing contextual associations during a set of initial presentations may promote encoding a leitmotif as a long-term memory trace. Boltz et al. (1991) examines the effect of background music on remembering filmed (visual) events, and observes that the music’s impact is dependent on its (affective) congruency with the visuals: “in situations where the mood of the music corresponds to the affective meaning of the scene, memory should be quite high . . . depending on the mood congruency and placement of music relative to a critical scene, music can enhance subsequent recall relative to situations where no music occurs at all” (p. 600). In a subsequent study, Boltz (2004) explores the inverse relationship, “whether visual information influences the perception and memory of music” (p. 43), and observes that “the affect and format of visual information” (p. 54) influences the cognition of music, particularly the cognition of melodies. The more emotionally and gripping the setting, therefore, the less we attend to generic framing elements; this sets up the conditions for salient (affective) elements to make a more memorable impression. 29. Much research suggests that leitmotifs significantly influence our attitudes toward particular ‘reallife’ scenarios outside of films, and that the impact of leitmotifs extends from in-theater to out-oftheater behavioral responses, particularly after exposure to highly effective horror or frightening films. Nosal et al. (2016) argue that “the music accompanying shark footage is nontrivial . . . many people trace their fear of sharks to the 1975 blockbuster Jaws, whose redolent soundtrack has become deeply rooted in popular culture . . . [evoking] haunting images of surfacing dorsal fins, swimmers’ legs underwater, and the histrionic combination of blood and bubbles” (p. 2). Cantor (2004) echoes these observations and expands the repertoire of films to explore the behavioral changes and lingering effects of leitmotifs. These lingering effects include sleep disturbances or anxiety in related waking life situations, such as “difficulty swimming after Jaws (in lakes and pools as well as the ocean); uneasiness around clowns, televisions, and trees after Poltergeist; avoidance of camping and the woods following The Blair Witch Project; and anxiety when home alone after Scream” (p. 283). While leitmotifs embedded in horror films are highly effective and clearest to unpack in terms of evolutionary psychology, leitmotifs embedded within westerns, romantic comedies, or sci-fi films may engender analogously strong psychophysiological responses.

7 ARCHETYPES

We have our tickets for the matinee, but we arrive a few minutes late. In our rush, we take no notice of the auditorium number—this is a massive multiplex cinema, replaying dozens of movie favorites from years past, all about to start. We trust, however, that our ears will guide us to the right auditorium. Walking past the first, we hear Christmas music, with the archetypical cowbells yet infused with sinister undertones—it must be Home Alone. Steps later, Spanish-sounding (Phrygian-mode) melodies on a flamenco guitar—most likely The Mask of Zorro. Then, Baroque flourishes on a harpsichord supporting a chamber string quartet’s contrapuntal gestures—Dangerous Liaisons plays here. Further ahead in the hallway, a short, pungent motif on a reverberant honky-tonk piano blending with sporadic timpani hits, a muted trumpet, and gunshots—likely a Spaghetti Western or a Tarantino movie borrowing the distinctive sound. Finally, as we hear synthesizer drones engendering futuristic sonic landscapes, we know we have found our auditorium—Blade Runner’s opening credits are rolling. Film music connotes distinct sociocultural spheres. It embeds musical archetypes that function as signs, guiding us through a film’s storyworld by revealing the ethnic or social background of characters, suggesting or supporting narrative developments, setting locale and time period, and even indicating the genre of films themselves. These signs are called musical ‘topics’—cultural constructs that transpire through a community of listeners’ musical discourse.1 Musical topics, however, are not permanent semiotic constructs. As cultural units of meaning, they emerge, change, and vanish with the constant flow and transformation of compositional and film-scoring practices.2 Therefore, in this chapter, we explore film music topics from two interdependent perspectives: first, from a synchronic perspective, describing their most common functions within a film at a single point in time from our vantage point today; and then, from a diachronic perspective, tracing a single topic’s development and transformation within U.S. films over nearly a century. Synchronic Perspective of Film Music Topics

A synchronic exploration of musical topics seeks to recognize their reception and signifying potential at a particular point in time. This perspective gives us access to four vital functions

DOI: 10.4324/9780429504457-8

Archetypes

101

of film music topics: to set locale and time period, to support character construction and development, to add subtext to the story, and to indicate genre. Locale and Time

Strategically placed in a film’s soundtrack, topics help set locale or time period, particularly by drawing on the connotative power of timbre. Distinct instrumental timbres function as topical markers to indicate precise geographic locations (guzheng for China, bandoneon for Argentina, ranat-ek for Thailand), precise location and time (distorted electric guitars for the 1980s U.S., harpsichord for 1700s Europe), or even somewhat undefined environments and times (synths and electronica for ‘somewhere in space, at some point in the future’). Whereas films in which the narrative unfolds within a single time and place may effectively introduce this topical function within the main titles, films in which the narrative navigates through disparate times or places may draw on this topical function within individual cues.3 Some topics help establish both time and place with precision. The main title music of Dangerous Liaisons features a harpsichord, a harp, and a small string ensemble, all performing in a Baroque style, situating the narrative within a well-defined time and place— eighteenth-century Europe. By introducing the archetypical dotted figure of French overtures, the music sets the story more precisely within the French royalty of the time.

FIGURE 7.1A

Dangerous Liaisons. Main title music. (Music by George Fenton.)

FIGURE 7.1B

Dangerous Liaisons. Main title sequence.

102

Archetypes

FIGURE 7.2A

The Last Emperor. Main title music. (Music by David Byrne, Ryuichi Sakamoto,

Cong Su.)

FIGURE 7.2B

The Last Emperor. Main title sequence.

In some cases, topics can be more suggestive of place than time. The main title theme of The Last Emperor, for example, introduces an ‘Asian pastoral’ topic. It features an erhu (a Chinese instrument akin to a violin) performing a pentatonic melody supported by taiko drums (a percussion instrument indigenous to much of Asia). The instrumentation and pitch configuration help place the narrative in the context of China. Nevertheless, because these instruments have been used in Asia for centuries, embedding such topical markers in the music helps foreground the place, rather than a specific time.

FIGURE 7.3A

Home Alone. Main title music. (Music by John Williams.)

FIGURE 7.3B

Home Alone. Main title sequence

.

Archetypes

103

Some topics indicate time, but instead of pointing to a time period, they suggest a season. The main title music in Home Alone, for instance, introduces a ‘Christmas’ topic to situate the narrative around late December of some undetermined year. The music opens with an innocent melody in the high register of a celesta outlining a major triad. However, as it adds the archetypical sleigh bells and triangle along with high-pitched woodwinds and strings, it swiftly turns to the parallel minor mode and surreptitiously slinks mischievous string glissandi and descending chromatic lines reminiscent of Tchaikovsky’s “Dance of the Sugar Plum Fairy”, evoking a ‘wicked Christmas’ topic. Here, the time of the year is more relevant than the specific year. Character Construction and Development

By supplying an additional layer to the identity developed by the dialogue, costume design, and setting, topics effectively aid in character construction, particularly when defining a character’s ethnicity or sociocultural background. Here, I use a single film, 15 Minutes, to distill the expressive power of topics in character construction. In the film, four characters from divergent sociocultural worlds cross paths: two ex-convicts from Eastern Europe, a laidback, old-school New York City detective, and a zealous young New York City fire marshal. In the film, when presenting these characters, the musical fabric weaves in archetypical topics to reveal vital facets of their identities. Emil and Oleg arrive in America from Eastern Europe. As they wait in line at customs and passport control, an ‘Eastern European’ topic emerges from the musical texture. This topic is characterized by the profuse use of the ‘Hungarian minor’ mode [D-E-F-G#-A-B♭C#-D], whose unusual structure contains four minor-second and two augmented-second intervals, and which, in combination with other musical markers (such as timbre and instrumental techniques), instill an exotic or ‘foreign’ sound unfamiliar to most Western listeners.

FIGURE 7.4A

15 Minutes. Main title music. (Music by Anthony Marinelli.)

FIGURE 7.4B

15 Minutes. Emil and Oleg arrive at JFK. [00:00:20]

104

Archetypes

FIGURE 7.5

15 Minutes. NYPD Detective Eddie Flemming sobering up. [00:04:30]

FIGURE 7.6

15 Minutes. NYC Fire Marshall Jordy Warsaw. [00:17:45]

In contrast, a ‘suave noir detective’ topic introduces world-wise NYPD Detective Eddie Flemming, first seen sobering up by submerging his head in ice water. The music, a smoothjazz quasi-improvised piece in Dorian mode reminiscent of Miles Davis’s “So What” and featuring a small ensemble comprising Hammond organ, muted trumpet, electric bass, and drums, pours a chilled, smooth, vintage veneer over Flemming’s character. Later, a ‘Miami Vice’-like topic introduces the young New York City fire marshal, Jordy Warsaw. The music features a metrically offset electric guitar riff over a looped beat-machine backdrop, imbuing Jordy with youthful urgency and dynamism. Throughout the film, although the music accompanying each character continues to infuse archetypical topical markers, it evolves, reflecting the characters’ journeys as they adjust to their new realities. Narrative Support and Subtext

Topics may also supply an additional narrative layer that depicts or expands on the connotations of the visuals or the dialogue. They do so by embedding markers we commonly associate with distinct settings (e.g., weddings, funerals) or affective conditions (e.g., sensual, energetic, humorous, relaxing). The ‘funeral march’ is a relatively stable topic, traceable centuries back to classical music.4 This topic, which features a minor mode, dotted rhythms, low register, slow-to-moderate pace, and duple meter, brings about associations that include death, mourning, and cemeteries. Introducing a ‘funeral march’ topic in The Three Musketeers (1948) quickly and effectively sets the stage for the event about to unfold. The three musketeers capture Lady de Winter and bring her to justice for the murder of the Duke of Buckingham. As she somberly walks toward her execution, the music introduces several key markers of the ‘funeral march’ topic.

Archetypes

FIGURE 7.7A

‘Funeral march’ topic in The Three Musketeers. (Music by Herbert Stothart.)

FIGURE 7.7B

The Three Musketeers. Lady de Winter walks toward her execution. [01:58:30]

105

In comedies and parodies, the music often over-intensifies the underlying affect by infusing a topic’s hyperbolic rendition, capturing a topic’s musical essence or most salient features, effectively using the soundtrack to evoke extra-musical associations. A scene from Meet the Parents, for example, portrays Greg Focker as a hero by introducing a version of the ‘heroic’ topic in the music—characterized by the use of trumpet and French horns, marching tempo, duple meter, periodic phrase structure, harmonic progressions based on diatonic triads in root position, profuse melodic intervallic motions in perfect 4ths and 5ths, and kettledrums or cymbals on selected downbeats.5 The ‘heroic’ topic in this scene, particularly its hyperbolic rendition, functions not as character construction but as character representation to support Greg’s short-lived heroic narrative—an equally epic downfall will soon follow his grand farce.6

FIGURE 7.8A

‘Heroic’ topic in Meet the Parents. (Music by Randy Newman.)

106

Archetypes

FIGURE 7.8B

Meet the Parents. Greg arrives victorious with the cat. [01:13:10]

Genre Indicator

Weaved through the musical fabric of main title themes, topics help filmmakers (and viewers) situate a film within a particular genre or sub-genre. Although main title themes are often extended pieces of music, by drawing on topics, they instantly convey the type of film about to unfold—sci-fi, horror, comedy, noir, or any other genre. This subsection explores the relatively stable ‘Spaghetti Western’ topic, which has maintained its semiotic currency for decades, outlasting the genre itself. The origin of the ‘Spaghetti Western’ topic can be traced back to Ennio Morricone’s film scores and his collaborations with director Sergio Leone. In particular, his main theme for The Good, the Bad and the Ugly includes topic-defining markers: galloping rhythms rendered in the percussion or strummed guitars, Aeolian mode, whistling or animal howls (e.g., the ‘coyote’ motif), short phrases with relatively lengthy pauses, ‘twangy’ guitar, harmonica, bass ocarina, soprano recorder, and strings or choir background textures.7 In addition, sonic markers providing a stylized rendition of on-screen environments—such as long reverbs suggestive of vast, empty, open spaces, and close microphone techniques matching the extreme closeups—helped define the archetypical sound of the ‘Spaghetti Western’ topic.

FIGURE 7.9A

The Good, the Bad and the Ugly. Main title music. (Music by Ennio Morricone.)

FIGURE 7.9B

The Good, the Bad and the Ugly. Main title sequence.

Archetypes

FIGURE 7.10

107

The Mandalorian, Season 1, Episode 1. Opening sequence.

The main title theme from The Mandalorian borrows a great number of markers of the ‘Spaghetti Western’ topic: it opens with a solo ocarina ‘howling’ in its mid register and a short melodic figure reminiscent of The Good, the Bad and the Ugly, all immersed within a reverb and echo suggestive of a large space. Moments later, a synthesized animal cry and a bell-like hit preface galloping rhythms in the drums, which serve as accompaniment to an Aeolian-mode call-and-response melodic exchange between acoustic guitar and low brass. Toward the center part of the theme, as the galloping accompaniment continues, a brief interlude introduces a whistling timbre resembling a Theremin, supplying a futuristic aura. Although this TV show is set in the Star Wars storyworld, ostensibly science fiction, these topical markers narrow and (re)direct our expectations in a different direction—we will most likely expect a lone, possibly flawed (anti-)hero with an idiosyncratic moral code. As a result, by weaving archetypical topical markers through the fabric of the main theme, the filmmakers firmly situate The Mandalorian within the Spaghetti Western genre’s legacy. Combined Topical Functions

The functional categories delineated so far are not mutually exclusive; in fact, topics are most effective when simultaneously serving various functions. In Who Framed Roger Rabbit?, an animated film noir, the ‘sexophone’ topic performs multiple functions. This topic, musically characterized by a short and sultry saxophone riff, originated within film noir’s smooth-jazz sound world, chiefly when introducing the femme fatale. This topic has outlasted the noir genre itself, continuing to permeate other film genres to indicate the presence of a sexually suggestive character or a moment of sexual magnetism between characters. In the film, a sensual saxophone riff consistently signals Roger Rabbit’s human-toon wife, Jessica Rabbit. She is a ‘real’ femme fatale—bodacious, seductive, mysterious, duplicitous—who claims, “I’m not bad, I’m just drawn that way.” The ‘sexophone’ topic in these scenes thus indicates the film’s genre, defines one of its characters, and supports moments of physical attraction between characters in the narrative.

FIGURE 7.11

Who Framed Roger Rabbit? ‘Sexophone’ topic. [01:43:50]

108

Archetypes

Diachronic Perspective

While from a synchronic perspective topics seem stable and permanent units of meaning— featuring little or no deviation in their archetypical markers—a diachronic perspective reveals that they are fluid constructs: new topics emerge while existing ones change their markers, fuse with other existing topics, or vanish altogether. In this second part of the chapter, an abridged diachronic analysis of the ‘superhero’ topic permeating through a wide range of superhero films’ main title themes reveals significant changes in the topic’s markers over time.8 The ‘Superhero’ Topic

Early in the twentieth century, many superhero characters, including Superman, Batman, and Wonder Woman, arrived within (the soundless media of) comic strips. With the advent of new technologies, these characters made inroads into the emerging media of television and cinema, which, over time, have become dominant genres with their own iconography and their own fluid yet broadly recognizable music topics.9 Because present-day audiences encounter superhero characters primarily through the (music-filled) movies, superheroes are inextricably linked with their accompanying musical themes. Superhero themes emerge from and help establish the ‘superhero’ topic—changes to the topic transpire when a superhero theme takes on additional musical gestures to reflect distinctive facets of a superhero. Changes in a single film lead to topical representation of similar characters in other films, and in some exceptional cases, themes deviate significantly from the norm and help define radically new trends in this musical topic.10 Over time, new

FIGURE 7.12A

Captain Blood (1935). Main title music. (Music by Erich W. Korngold.)

FIGURE 7.12B

Captain Blood (1935). Main title sequence.

Archetypes

109

FIGURE 7.13A

Superman (1941). Main title music. (Music by Winston Sharples & Sammy Timberg.)

FIGURE 7.13B

Superman (1941). Main title sequence.

gestures and figures gain cultural currency, becoming the new signifiers of the ‘superhero’ topic. In the 1930s and 1940s, while searching for musical gestures that would support and help construct the superhero persona, film composers drew on established musical traditions (opera, concert music). They infused superhero themes with the post-romantic sound of epic orchestral forces, sweeping melodic gestures, and other archetypical features of the previously established ‘military’ topic—duple meter, brisk tempo, major mode, brass and percussion timbres, diatonic progressions, triadic chordal structures, periodic phrases. The main theme for Captain Blood (1935), for example, emerges as an early instance of a topically defining theme, featuring a major-mode trumpet fanfare with profuse close-voiced triadic structures and stable periodic phrasing typical of the ‘military’ topic. (See Figure 7.12A.) The theme for the Superman (1941) animated cartoon series adopts the trumpet fanfare and triadic gestures associated with the ‘military’ and (at the time) ‘superhero’ topic but detours into non-diatonic regions via a chromatic ♭VII harmony, which one may read as alluding to the ‘superhuman’ powers of the character, a characteristic absent in ‘human’ heroes.11 The music thus absorbs and crystalizes the superpowers of the characters it represents so as to transcend the familiar diatonic boundaries. (See Figure 7.13A.) In the 1960s, the Batman TV series theme illustrates a marked shift away from the epic musical flair characteristic of earlier superhero screen renditions. Because the series fitted seamlessly within a camp film noir narrative—with intricate plots that revolved around assaults, thefts, murders, and moral corruption—it seems suitable that its main theme

110

Archetypes

FIGURE 7.14A

Batman TV series (1960s). Main title music. (Music by Neal Hefti.)

FIGURE 7.14B

Batman TV series (1960s). Main title sequence.

borrowed elements from noir soundtracks, especially those by Henry Mancini (Peter Gunn, The Pink Panther), Count Basie (M Squad), and Monty Norman (Dr. No, the James Bond theme).12 After a fleeting ‘fanfare’ gesture in the trumpets, reminiscent of previous superhero themes, the Batman TV series theme outlines a twelve-bar blues progression (with its distinctive extended harmonic language), features a small-band ensemble highlighting a twangy guitar performing a swirling riff around #4ˆ, introduces vocals replicating the guitar riff with the catchy ‘na na na na na na na na . . . Batman!’, and embeds sonic markers that emulate the visual expletives emblematic of comic books—‘SOCK!’, ‘POW!’, ‘ZOK!’. Here the quasi-noir sound and the sonic markers contribute to the TV show’s tongue-in-cheek rendition of its comic book counterpart. (See Figure 7.14A.) The shift away from the grand symphonic sound extended well into the 1970s, with the main theme for the Wonder Woman TV film (and later the TV series) following suit—featuring a Boogie-Woogie accompaniment, big-band instrumentation, and pop vocals.13 Nevertheless, a brief reference to the prior tradition of superhero themes remains in the form of a short brass fanfare. In the film, a U.S. Air Force plane crashes on an uncharted island inhabited by Amazon women, one of whom, Diana, brings the pilot back to the U.S. and helps battle the Third Reich. Diana strives to adapt to the American culture, wearing ordinary clothes and getting conventional jobs (e.g., nurse, secretary). In fact, when she takes on a superhero role (as Wonder Woman), she reveals her skimpy, star-spangled armor and her special powers. By appropriating a big-band, iconic American sound, Wonder Woman’s main theme helps her fit in and become an American cultural icon. (See Figure 7.15A.)

Archetypes

FIGURE 7.15A

Wonder Woman TV series (1970s). Main title music. (Music by Charles Fox.)

FIGURE 7.15B

Wonder Woman TV series (1970s). Main title sequence.

111

The release of Superman: The Movie in 1978 marked a return to orchestral superhero themes infused with the ‘military’ topic, including markers such as major key, brass timbres, duple meters, and fanfare-like gestures.14 Its harmonic stability notwithstanding, and just like its 1941 counterpart, the theme for Superman (1978) detours into non-diatonic harmonic regions. Here, a non-diatonic harmonic gesture now known as ‘Aeolian cadence’ (♭VI ® ♭VII ® I) becomes intricately linked to the character’s heroic actions. Since then, this musical gesture has been highly influential, emerging as a distinct marker of the ‘superhero’ topic, reliably found in the music of more recent films, such as in Captain America (2011).15

FIGURE 7.16A

Superman (1978). Main title music. (Music by John Williams.)

112

Archetypes

FIGURE 7.16B

FIGURE 7.17

Superman (1978). Main title sequence

.

Captain America (2011). Main-on-end title sequence

.

Batman (1989) once more brought a new bent to the musical signatures of the superhero topic. Although Batman’s main theme maintained the brass-heavy orchestration and the militaristic gestures, it also introduced a darkly minor mode and a disquieting melodic gesture [♭6ˆ ® 5ˆ ® #4ˆ] carrying devious and tragic resonances.16 Additionally, the theme often interweaves the ‘uncanny’ topic—characterized by oscillating chromatic mediants, particularly including minor chords—bestowing the character with an ominous aura.17 The film begins on a dreary night in Gotham City as the Wayne family exits a movie theater and ventures down a poorly lit alleyway. Emerging from the shadows, a man wielding a gun demands their valuables, and the Waynes submit without resistance. A high-angle shot captures a fleeting shadow. To confirm Batman’s looming presence, the music swells with his signature theme, blending a chromatic mediant relationship (Cm ® A♭m) to evoke dark and uncanny resonances synonymous with Batman’s nocturnal persona.

FIGURE 7.18A

First appearance of Batman’s theme. (Music by Danny Elfman.)

Archetypes

FIGURE 7.18B

113

Batman (1989). First suggestion of Batman’s presence. [00:04:30]

As we consider more recent superhero films, it becomes nearly impossible to discern which soundtracks will influence current and future ‘superhero’ topics. However, a cursory corpus study of recent influential superhero themes reveals musical markers with the potential to establish new trends: symphonic timbres blended with electronica or unusual timbres, asymmetric meters, and minor modalities infused with exotic chromaticism. For instance, the theme for the most recent Wonder Woman franchise (2017, 2020) foregrounds distorted electric guitars accompanied by strings and brass, projects an asymmetric septuple meter (2 + 2+2 + 1), and introduces #4ˆ (instead of its diatonic counterpart) as its melodic focal point. These musical signatures reflect on the narrative and help construct and support Wonder Woman’s persona—the electric guitars infuse a retro feel that helps situate the narrative in the 1980s, and the asymmetric meters along with the melodic emphasis on #4ˆ colors Wonder Woman’s character with metrical and tonal qualia that evoke restlessness and curiosity.18

FIGURE 7.19A

Wonder Woman (2017). Main-on-end title music. (Music by Rupert GregsonWilliams.)

FIGURE 7.19B

Wonder Woman (2017). Main-on-end title sequence.

114

Archetypes

Coda

Just as cultures rely on archetypes to define their identities and enable communication, films strongly depend on musical topics—culturally shared musical symbols to supply information about the film, the characters, the setting, or the narrative. In this chapter, we explored film music topics from two complementary perspectives: synchronic and diachronic. A synchronic perspective allowed us to distill the phonetic and syntactical structures on the music’s surface that denote specific extramusical meanings. In turn, a diachronic perspective allowed us to unearth old topics and trace their influence in (trans)forming new ones, while reminding us of a topic’s (im)permanence as a unit of meaning permeating filmgoers’ cultures.19 As we walk out of the multiplex, more recent films are about to begin. We first hear a catchy pentatonic motif on exotic zither-sounding instruments (yangqin and guzheng) punctuated with emphatic percussion (tanggu, paigu, and bangu drums)—musical markers evoking an ‘Asian’ topic. Then, symphonic forces join in, with proud strings on a powerfully tenacious minor-mode melody and heroic brass outlining triadic progressions that project an unyielding harmonic stability delicately decorated with exotic chromaticism—musical markers denoting a contemporary ‘superhero’ topic. Such an exquisite fusion of topical ethos suggests that Shang-Chi and the Legend of the Ten Rings plays in this auditorium. We can hardly wait to come back for more! Notes 1. As signs, musical topics are akin to leitmotifs. However, whereas a leitmotif’s signified remains exclusively within a film’s storyworld, a topic’s signified resides outside a film’s storyworld, in the collective subconscious shaped by soundtracks to multiple films. 2. Within a community of listeners sharing common musical codes, topics are effective means for conveying meaning; outside of such a community of listeners, however, topics may lose their semiotic currency and even go unnoticed. For an in-depth discussion of the notion of archetypes, particularly musical topics, see Appendix VI. 3. Examples of the latter include Back to the Future (1985), Jumper (2008), Hot Tub Time Machine (2010), Looper (2012), and About Time (2013). 4. Particularly relevant examples include Chopin’s Piano Sonata No. 2 (third movement), Beethoven’s Eroica Symphony (second movement), and Mendelssohn’s Song Without Words Op. 62, No. 3. 5. Although the sociocultural associations corresponding to this topic seem grounded in a symbolic relationship (i.e., based on convention), these rest on both iconic (i.e., through resemblance) and indexical (i.e., through proximity) relationships—for instance, the duple meter functions iconically, mapping the physicality of marching, while the sound of kettledrums and trumpets functions indexically, indicating early battlefields. 6. See Chapter 9 for an in-depth discussion of this scene through the lens of categorization. 7. Leinberger (2004) offers a fascinating, in-depth study of Morricone’s score for The Good, the Bad, and the Ugly. 8. Although we attend here to the diachronic fluidity of the ‘superhero’ topic by focusing on new markers introduced over time, one may understand much about this and other topics by attending to their invariant markers—that is, attending to the consistent musical features across diachronically different versions of a topic. Additionally, one may construct a hierarchy of topical stability by observing and computing degrees of diachronic difference—such a hierarchy may convey significant insights into the evolution of musical signs. 9. Buhler (2016) highlights the influential nature of music media in the realm of franchising, as it expands its reach to various platforms such as television, video games, books, amusement park rides, and websites.

Archetypes

115

10. Young (2013) identifies a blend of the ‘military’ and the ‘fantasia’ topics as the core sub-components of a ‘superhero’ topic and traces the origin of this blend to early sound film: signifiers of the military topic include march (duple meter) meters, major-mode (triadic) fanfares, and brass instrumentation, while signifiers of the ‘fantasia’ topic include harp glissandi and pantriadic chromaticism. In particular, pantriadic chromaticism seamlessly aligns with the elicited associations of the supernatural (and superheroes) by fostering a musical environment devoid of tonal forces via a “thorough negation of tonal norms of centricity, diatonicity, and functionality” (Cohn, 2012, p. xiv). 11. Young (2013) notes that “heroes such as Captain Blood, Robin Hood, and Zorro possess no extraordinary powers outside of prodigious skill with a sword—in short, they could exist in reality. However, heroes like Superman and Captain Marvel have abilities that are superhuman or godlike, demanding that they can only exist within a fantasy world” (p. 111). 12. The Batman character aligns with the three types of film noir ‘heroes’ identified in the literature: the ‘seeker-hero’, whose “investigation takes the form of a quest into a dangerous and threatening world: the noir world” (Walker, 1992, p. 10); the ‘victim hero’, who is the “passive subject of investigation tested by threats to his masculinity and individual autonomy” (Mason, 2011, p. 138); and the ‘amnesiac-hero’, who “becomes a victim of a violent and hostile world and who lives in fear” (Walker, 1992, p. 15). 13. While the story for the first season of this series unfolded during World War II, the stories for subsequent seasons unfolded during the 1970s. Correspondingly, the lyrics of the TV series’ theme song, with its World War II references, were omitted after the first season. 14. Halfyard (2013) argues that “what gives [this] formula its heroic character is largely the harmonic stability created by the use of tonic-dominant harmony and corresponding melodic intervals such as prominent open fifths, alongside the energetic character of the march rhythms . . . The punchy, confident, militaristic scoring” (p. 172). 15. While in the Super Mario (1985) video game an Aeolian cadence indicates the successful completion of a task, the subsequent film, Super Mario Bros. (1993), embeds this harmonic gesture in its main title theme. The Aeolian cadence also often appears in songs associated with superhero films, such as Nickelback’s “Hero” for Spider-Man (2002) and Seal’s “Kiss from a Rose” for Batman Forever (1995). 16. Donnelly (1998) construes Batman’s theme as “pure Gothic melodrama, using a large, dark and Wagnerian orchestral sound” (p. 148), and Young (2013) reads it as a “[shift] from a strictly heroic style to one of the tragic heroic” (p. 105). Halfyard (2013) compares Batman’s and Superman’s themes and notes that Batman’s “minor-key theme draws on horror-movie musical tropes as much as superheroic ones, and the score substitutes constant harmonic slippage for Superman’s diatonic stability” (p. 175). 17. Bribitzer-Stull (2015) notes this chromatic-mediant oscillation in Richard Wagner’s Der Ring des Nibelungen, and suggests it denotes “mystery, dark magic, the eldritch, and the otherworldly” (p. 144). In tracing the affective impact of such chromatic-mediant oscillations, Lehman (2018) suggests that “the characteristic semitonal displacements of the pillars 1ˆ and 5ˆ of the home triad, pitches that are flayed outward in opposite directions, as though being tugged by invisible tonal tendrils of ill intent” (p. 101). Heine (2018) notes that this progression is so “strongly uncanny” that “only a sorcerer could conjure up such unnatural harmonies” (p. 122). Moreover, Huron (2006) conducted a series of experiments in which participants identified the qualia for chromatic mediants; the reported phenomenal responses (or qualia) for this chromatic mediant relationship include “mysterious, cheerless, somber, dark, tragic, despairing, death, depressed” (p. 273). Halfyard (2004) offers a fascinating in-depth study of Elfman’s score for Batman. 18. Huron (2006) notes that scale degrees (including chromatic ones) elicit uniquely different phenomenal responses, which we conceptualize using descriptors known as ‘qualia’. He reports that the raised subdominant (#4ˆ) evokes qualia that include “intentional, motivated . . . moderately anxious . . . curious about possibilities” (p. 145). 19. Bourne (2021) suggests a shift in directionality between concert and film music, proposing that contemporary listeners draw upon their familiarity with film music to understand Western art music.

8 ASSOCIATIONS

In The Conversation, the diegetic music surreptitiously infiltrates the narrative and subliminally informs our interpretations. After a tense day at work, Harry, his male colleagues, and their new female acquaintances drive to a warehouse for a spur-of-the-moment party. Not comfortable in large gatherings, Harry seeks comfort in Meredith, a young woman he just met. Meredith cuddles close to Harry as she dances him away from the crowd. As they drift to a secluded area, an instrumental version of Duke Ellington’s “Sophisticated Lady” softly reverberates through the warehouse. They dance intimately for a while. At ease, Harry and Meredith open up about their lives as the song continues distantly in the background. During this moment of tender intimacy, the diegetic music reveals facets of her personality: the song’s associations impinge upon our (and possibly Harry’s) perception of Meredith—a sophisticated lady, “smoking, drinking, never thinking of tomorrow”. Film music draws from deep within our personal associations to embed subtextual commentary, eliciting uniquely individualized readings of a single scene or an entire film. Conceptual Blending, a framework developed by Fauconnier and Turner (2002), is exceptionally suitable for fleshing out potential projections between the music’s associations and a film’s dramatic development that elicit those readings. In this chapter, we draw on this framework’s multiple-space model to explore instances in which existing songs’ or concert pieces’ contextual associations (even if distinctly subjective) contribute to a film’s meaning with subtexts that inform our interpretations. Songs

Filmmakers recognize the potential of introducing songs to supply an additional layer of meaning that supports and comments on the main narrative. Often, songs (even instrumental versions) from different styles, eras, and languages strategically align with the narrative events. At these moments in a film, we draw onto a song’s associations (elicited by its lyrics or title) and project these onto the film’s narrative to construct an interpretation. Lars and the Real Girl explores the nature and reality of love. Lars, a quirky young man conflicted by feelings of love and loss, develops a romantic (non-sexual) bond with Bianca,

DOI: 10.4324/9780429504457-9

Associations

FIGURE 8.1A

Lars and the Real Girl. Lars serenades his beloved. [00:42:40]

FIGURE 8.1B

Conceptual blending in Lars and the Real Girl.

117

a realistic sex doll, forging an imaginary relationship out of love and loneliness that thrives on pseudo-dialogue and a pretense of tolerance. In the Midwest woodlands, placidly resting in a treehouse, Lars sings Nat King Cole’s classic song “L-O-V-E” to Bianca: “L is for the way you look at me, O is for the only one I see, V is very very extraordinary, E is even more than anyone that you adore can . . .” He continues singing, “Love is all that I can give to you, Love is more than just a game for two, Two in love can make it”, and abruptly stops to suggest, “You can watch me chop wood, too. I’m really good at it.”1 As the conceptual blend in Figure 8.1 outlines, the associations brought about by the song’s lyrics inform our interpretation of the film.2 The first stanza reveals Lars’s delusional version of love, one that flouts the reality of Bianca’s inanimacy—by switching between chest voice and falsetto, he simulates a feminine timbre, projecting onto Bianca a (false) voice with which she appears to join him in a duet. In turn, Lars’s singing only the first three lines of the second stanza helps delineate the film’s narrative trajectory: love is all that Lars can give to Bianca; his love is more than just a game and one that extends beyond the two of them to involve Lars’s family and the entire town; and these feelings, however delusional, help Lars make it, as everyone around him grows more human by accepting Bianca, helping Lars overcome his trauma. Compilation scores saturate the soundtrack with pre-existing popular songs. This practice emerged as a marketing strategy, initially placing a catchy feature song at the film’s beginning or end.3 Then, during the 1980s, the financially advantageous idea of selling a movie and a soundtrack prompted filmmakers to introduce many popular songs as part of the non-diegetic (rather than diegetic) soundtrack, songs that (often) bore little or no relation to

118

Associations

the events in the film.4 Nevertheless, this practice opened new dimensions in the general conception of film music, a practice that, in the hands of thoughtful music supervisors, enhances a film’s story through fitting and ingenious song placement. Fools Rush In is a romantic comedy in which dozens of songs from different cultural backgrounds heighten the marked differences in the protagonists’ cultural backgrounds. Alex, a New York nightclub designer, meets Isabel, a beautiful young Mexican photographer. Their one-night stand results in Isabel’s unplanned pregnancy. In a scene toward the beginning, Isabel unexpectedly reappears at Alex’s doorsteps, three months after their romantic encounter. Alex arrives home and exclaims in surprise, “Isabel!” She concedes, “You remembered. Well, I didn’t know what to do. I never did anything like that before, going home with someone I don’t know.” In an effort to comfort her, Alex adds in a scattered voice, “Hey, you and me, both. It was just one of those great, phenomenal, spontaneous things.” An awkward silence sets in. To make conversation, Alex asks, “So, how you been?” to which Isabel unhesitatingly responds, “Pregnant.” Alex is stunned by the news. His reaction upsets Isabel, who drives away in a huff. As Alex follows her, the song “Para Dónde Vas?” [“Where Are You Going?”] emerges in the soundtrack. At a surface level, the lyrics speak Alex’s mind as he follows Isabel—we may read “Para dónde vas, muchacha?” as “Where are you going, Isabel?” At a deeper level, the conceptual blend in Figure 8.2 illustrates how this Spanish song can be heard

FIGURE 8.2A

Fools Rush In. The Iguanas’ “Para Dónde Vas?” begins to play. [00:20:30]

FIGURE 8.2B

Conceptual blending in Fools Rush In.

Associations

119

as Alex attempting to speak to Isabel in her language, acknowledging her culture and his willingness to engage with it, seeking to bridge the gap between them and trying to repair the damage caused by his initial reaction. Whereas in the preceding examples the soundtrack foregrounds the lyrics, films often feature instrumental versions of songs with renditions that exclude the lyrics. In such cases, the potential projections we form between the music and the narrative rest on our recognition of the song and prior knowledge of its lyrics or title. In the provocative revenge thriller Promising Young Woman, numerous songs, some transformed beyond recognition, supply subtexts that (re)define the characters. Cassie, a brilliant medical student, drops out of school as she struggles emotionally and psychologically after her best friend’s rape and subsequent suicide. By day, she works at a coffee shop; by night, she reclaims her agency, asserting her power as a woman by adopting an alter ego to confront male predators. Toward the film’s end, Cassie plans to infiltrate the bachelor party of the men who had assaulted her friend. It is early in the evening. Cassie pulls up on a deserted dirt road in the woods and adds finishing touches to her makeup. Heavy black mascara and eyeliner magnify her dark eyes against the bleach-blonde wig, framed by large dangling silver earrings and a nurse’s cap. She is barely recognizable. In the music, an exotic, tense, ill-fated melodic figure outlining a tritone (F#, D, E♭, C) brings veiled yet lethal undertones. She steps out of the car, tosses the plate into the shrubs, grabs a nurse’s bag and red high-heel shoes from the trunk, and slowly strides down the road and uphill toward an isolated cabin in the woods. Now, the music’s minor mode, sinking string sweeps, and heavy pace suggest a somber march topic, escorting her to the danger zone, foreshadowing her fate. As we entrain to the music’s meter and subvocalize its deviant glissandi, we realize this is Britney Spears’s “Toxic” in disguise.5 Cassie arrives at the cabin, puts on her high heels, and rings the bell—she is about to penetrate a dangerous space. On the downbeat, one guy opens the door to the testosterone-filled cabin and announces, “Yes! The doctor is in the house!”6 This visual and musical foreplay reveals a cluster of discourses informed by the song’s title and its arrangement, mapping a plethora of associations and attributes onto the film’s narrative: the disguised song underscores Cassie’s transformation into her alter ego to infiltrate the bachelor party disguised as a stripper, yet her reaction to the male threat manifests itself through an augmented sexual agency that becomes just as deviant, just as toxic.

FIGURE 8.3A

Promising Young Woman. Cassie arrives at the cabin. [01:22:15]

120

Associations

FIGURE 8.3B

Conceptual blending in Promising Young Woman.

Concert Pieces

Selective projections between the music and narrative spaces prompt us to construct inferences, arguments, and interpretations. Such selective projections extend beyond those suggested by the lyrics: a piece’s program (preconceived narrative) or compositional milieu (place of composition, era, style) allows directors to convey vital information or supply a subtext critical to our understanding of a film’s narrative. The plot of the futuristic action thriller Minority Report revolves around a precrime unit in the police force. With the aid of three precogs (i.e., psychics), the police gather visual information and preempt future violent crimes. John Anderton, Chief of the Precrime Unit, enters the analytical room and takes off his coat. Jad, the main dispatcher, briefs Anderton, “We got a white male, about five-eight, approximately one-forty.” Anderton prepares to manipulate the visuals transmitted by the precogs to find information that would reveal the time and place of the crime and the offender’s identity. He puts on his finger gloves and inserts a disc into a slot. Schubert’s Unfinished Symphony begins to play. Jad shares, “I love this part”, as he watches Anderton’s conductor-like technique for parsing information, ordering the scattered images by skillfully placing some in the foreground, some in the background. Anderton zooms in onto one image to get a clearer picture of the gunman—it is Anderton himself. He abruptly stops the session; the music stops too. Introducing Schubert’s Unfinished Symphony prompts us to draw a metaphorical parallel between the music and the narrative—as its name implies, Schubert never completed this piece, which foreshadows a critical plot point later in the film. The conceptual blend in Figure 8.4 reveals the link between the music’s connotations with the plot: at the surface level, his skillful ‘conducting’ will help the police force prevent crimes from coming to fruition; and at a deeper level, within the broader context of the film, Schubert’s Unfinished Symphony, a famously incomplete piece from the distant past, may be heard as a reflection of Anderton’s grappling with unresolved trauma from his past, particularly his failed attempts to prevent the abduction of his son.

Associations

FIGURE 8.4A

Minority Report. Anderton manipulates information. [00:38:00]

FIGURE 8.4B

Conceptual blending in Minority Report.

121

Often, when a piece belongs to an opera or dance suite, the associated narrative or program of the larger work impinges upon our interpretation. Lord of War offers a fictional yet incredibly realistic glimpse into the end of the Cold War and the emergence of worldwide terrorism. The film follows smuggler Yuri Orlov unscrupulously supplying illegal weapons to emerging world powers. He titles himself a ‘Lord of War’. In an immorally sublime moment, Yuri admires an AK-47. To the mesmerizing sounds of Tchaikovsky’s “Swan Lake”, the camera zooms into and dances around the weapon—its chromed barrel, its elegant 30-round curved magazine. The correlations between the narrative and the music extend beyond the Russian origin of the machine gun and the musical piece, juxtaposing beauty and evil, disguising nefariousness in refinement. In Tchaikovsky’s ballet, Prince Siegfried falls in love with Odette, a young woman cast under the spell of an evil sorcerer who turned her into a beautiful white swan. To break the spell, Siegfried must pledge his love for Odette. But one night, at a social function at Siegfried’s castle, the magician appears in disguise with his wicked daughter Odile, who looks much like Odette yet wears a black dress. Deceived by the similarity, Siegfried mistakenly swears eternal love to Odile. The character associations triggered by Tchaikovsky’s work allow us to understand the scene in a new light. The conceptual blend in Figure 8.5 reveals the link between the music’s connotations and the plot: ‘Lord of War’ becomes mesmerized by the deadly black AK-47 machine gun.

122

Associations

FIGURE 8.5A

Lord of War. Yuri inspects an AK-47. [00:44:45]

FIGURE 8.5B

Conceptual blending in Lord of War.

In some unique cases, the associations that stem from a piece’s formal structure inform our interpretations of an entire film’s narrative structure. In the political thriller The Lives of Others, a piece titled “Sonata for a Good Man” becomes an important plot point. A “sonata” form in music provides the composer with a large-scale design with dramatic potential; it contains the three sections characteristic of rhetorical structures: exposition, development, and (transformed) recapitulation.7 A scene early in the film shows Georg and his fiancé Christa sharing some intimate moments after his birthday party. He briefly looks at one present, a score of the “Sonata for a Good Man”. They are not aware, however, that Hauptmann, a GDR officer codenamed HGW XX/7, is keeping them under close surveillance. As Hauptmann observes the day-to-day life of Georg and the antiGDR group, he begins to question his own ideology. As the story unfolds, he embarks on a radical transformation, sympathizing more and more with the anti-GDR group. The last scene of the film shows a profoundly changed Hauptmann. He no longer works as a secret agent and no longer shares the GDR’s ideology. Moreover, the dedication of Georg’s novel to Hauptmann suggests a complete disengagement from his previous GDR persona and his embracing of a greater concern for the human condition. The plot from The Lives of Others thus closely maps a sonata’s formal structure: it presents two main characters with contrasting ideologies; subsequently, it sets up a confrontation and struggle between these two ideologies; and ultimately, it presents once more the two main characters, but one has undergone a radical transformation, assimilating the values and perspectives of the other.

Associations

123

FIGURE 8.6A

The Lives of Others. Georg unwraps the score for the “Sonata for a Good Man”. [00:34:20]

FIGURE 8.6B

The Lives of Others. Hauptmann’s transformation. [00:53:00]

FIGURE 8.6C

The Lives of Others. Georg’s book, “Sonata for a Good Man”, catches Hauptmann’s attention. [02:11:30]

FIGURE 8.6D

Conceptual blending in The Lives of Others.

124

Associations

Coda

During a film, a network of musical, (con)textual, and personal associations influences our reading of a scene. Conceptual Blending offers a robust analytical framework to explore the projections we construct between the music and other cinematic domains, revealing (and reconstructing) the cognitive processes that prompt us to generate meanings and interpretations about music’s place within a film. In instances where the music does not offer new information but rather echoes the other input spaces, blending allows us to focus on the elements present in the input spaces by accentuating subtle differences.8 The Conversation comments on the emerging social tensions between the public and the private, articulating a rendition of a collective reality by extrapolating the inner psychological journey of an individual. Befitting the film’s themes of secrecy and wiretapping, the soundtrack echoes the film’s narrative without perceptible authorial intrusion: while the plot portrays a transgression of the protagonist’s private space, the soundtrack maps this transgression onto the aural boundaries, tapping into our personal, private musical associations. Notes 1. Lars halts his singing halfway through the stanza. He omits the last two stanzas of the chorus, “Take my heart and please don’t break it, Love was made for me and you”, which deviate from the film’s primary plot points—these lines point to the very agency that Bianca lacks, suggesting that love was not intended to include inanimate objects. 2. In this example, the vital relationship of ‘disanalogy’ links corresponding elements across input spaces. Therefore, running the blend highlights not the similarities, but the differences. For an indepth discussion of Conceptual Blending, see Appendix VII. 3. Addressing that trend, Bazelon (1975) mentions that “it does not seem to matter that the theme tunes have little relevance to the film’s dramatic context . . . . Usually placed at the beginning as a title song but occasionally at the end . . . the songs cash in on today’s fast-changing market, ostensibly giving pictures a gilt-edged frame of catchiness” (p. 30). 4. Criticism has been directed not only at the obvious financial motivations of a compilation score but also at its artistic function within the film. Bell (1994) notes that songs are “being purchased and placed in films, not for artistic reasons, but because they might sell more soundtrack records/CDs” (p. 66), and warns us that “the incorrect use of songs endangers the cohesiveness of film art. Instead of a two-hour dramatic statement, motion pictures often become bits of plot interspersed between MTV-like music videos” (p. 67). 5. Entraining to Cassie’s actions (by entraining to the music’s meter) affords us an opportunity to embody her steady (yet ill-omened) march and further identify with her. 6. At the cadence point, the final chord (the unstable dominant-seventh of the minor mode) opens up the space for the events about to unfold. Some listeners may hear the sound of the opening door as a downbeat that brings a level of resolution to the music via a sound design transference, one that transfers the agency from Cassie to the men in the house. 7. In the exposition, the music presents two contrasting themes—the first in the home key and the second in a contrasting key. In the development, it musically explores and confronts these themes. In the recapitulation, the music once more presents both themes, but the second one, now in the home key, has undergone a transformation to conform to, and arguably assimilate, the qualities of the first theme. 8. Elsewhere (Chattah, 2008), I propose a preliminary model for analyzing irony in film music based on a modified version of Fauconnier and Turner’s framework. Within this model, interpretations of irony emerge in a blended conceptual space and stem from cross-domain projections based on vial relations of ‘incongruity’ or ‘opposition’ between input spaces. However, this model presents a critical limitation: its inability to parse out various types of irony or to distinguish tropes closely related to irony. Therefore, in Chapter 9, I apply models of categorization to disentangle the tropes emerging from the deliberate placement of incongruous music in a film.

9 CATEGORIZATION

In Terminator 2: Judgment Day, the music often supplies contradictory subtexts. Strange lightning forms a circular opening in the sky, and a flare of light materializes as a massive, sculpted, naked body. The Terminator has arrived. It strides impassively into a diner, swiveling its head with its characteristic emotionless gaze. Customers freeze. It approaches a rough-looking biker and orders, “I need your clothes, your boots, and your motorcycle.” As it steps out of the diner, now fully clothed in black leather and heavy boots, George Thorogood and The Destroyers’ “Bad to the Bone” roars in the soundtrack. We recognize that the song functions as an off-screen narrator revealing facets of the character, yet we have seen the film already and know the Terminator is not human, has no bones, and is not the bad one here! It is so ironic . . . But is it? Film music conveys messages that support, complement, or negate the visuals or the dialogue. When confronted with conflicting, ambiguous, or incompatible meanings between the music and other cinematic domains, we generally resort to a chain of inferences suggesting ‘irony’. Indeed, a key component in irony is the presence of incongruity; but a detailed examination of these film moments reveals that tropes closely related to irony (such as parody, satire, or paradox) may instead be at play. In this chapter, we draw on probabilistic models of categorization and the notion of second-order inferences to construct a framework for reevaluating our interpretations of incongruity in film. Categorization of Irony and Related Tropes

Our interpretations of irony and related tropes are contingent upon the category structures we have formed of these rhetorical devices. Distilling the process whereby we arrive at interpretations of irony and related tropes, particularly within films, prompts us to deconstruct these category structures and explore the possible mechanisms that led to their construction.1 A probabilistic approach to categorization provides a flexible framework that allows us to include first- and second-order interpretations as attributes, assign weighted values to these attributes, and even identify nested or inter-attribute relationships.2 The categorization process begins with our constructing first-order interpretations based on textual attributes. These attributes include perceptually salient or relevant characteristics, DOI: 10.4324/9780429504457-10

126

Categorization

FIGURE 9.1

First-order interpretations as attributes and their corresponding values.

such as incongruity (conflicting presentation between the music and any other conceptual domain), intertextuality (borrowing from another film or another genre), hyperbole (out-of-context exaggeration), grotesque (distortion exposing peculiar characteristics), and cinematic anomaly (divergence from established cinematic conventions).3 First-order interpretations are also contingent upon the attributes’ values, which specify an attribute’s location within the cinematic text—for instance, “hyperbole” may be located in the music, the visuals, or the dialogue. Because we find traces of these attributes in the cinematic text, first-order interpretations seem objective; but forming first-order interpretations is also contingent upon extra-textual contexts, the discursive practices of interpretative communities, and the knowledge of culturally established aesthetic and cinematic conventions. In evaluating first-order interpretations (e.g., degree of hyperbole, nature of perceived anomalies), an additional set of attributes emerges as second-order interpretations: reversal (indicating a reassessment of semantic value that results in the opposite effect), critical appraisal (suggesting a target of judgment), focal shift (highlighting a perspective), and homage (celebrating an extra-opus text). These second-order interpretations are again cast as attributes, each with their corresponding values—for instance, “focal shift” may highlight the perspective of a character, the audience, or an unseen narrator. In contrast to attributes related to first-order interpretations, attributes stemming from second-order interpretations are not in the text, but in the audience’s mind.4 Ultimately, both first- and second-order interpretations combine to form a list of attributes, which are weighted when categorizing film examples.5 Exploring examples and identifying

Categorization

FIGURE 9.2

127

Second-order interpretations as attributes and their corresponding values.

the attributes that contribute to forming interpretations of irony and related tropes reveal near-prototypical exemplars begin to emerge.6 In assigning labels to categories, I resort here to preconceived notions developed outside of cinema and include tropes that suggest irony or that we commonly associate with (or mistake for) irony:7 • Irony proper ○ Structural (Situational) Irony: Expectations deviating from the state of affairs. ○ Dramatic (Tragic) Irony: Character’s fate revealed to the audience but unknown to the character. ○ Socratic Irony: Message of praise to imply blame, or of blame to imply praise. • Tropes related to irony ○ Sarcasm: Disguised contempt or criticism toward an individual. ○ Satire: Disguised contempt or criticism toward a contextual situation or social dynamics. ○ Parody: Exaggeration or decontextualization (of a style or genre) for comic effect.8 • Distantly related tropes commonly mistaken for irony ○ Quotation: Inter-textual reference. ○ Paradox: Self-contradictory statement. ○ Lie: Untruthful and deceiving statement.

128

Categorization

Irony and Related Tropes in Film Music

Although contextual expectations circumscribe the use of musical irony (and related tropes) to specific film genres—that is, we expect to encounter instances of irony and related tropes in comedies and parodies rather than in film noir or westerns—given the right conditions, irony may manifest itself in any genre. The expressive potential of all cinematographic facets, including music, bestows films with the rhetorical power to evoke a multitude of tropes; however, finding examples where most readers will agree upon an interpretation is no small feat.9 I trust, however, that most readers will agree with my suggestions regarding the rhetorical tropes the examples selected bring to mind. Irony is only effective when grasped by the viewer. When cinematic cues are not sufficiently salient, explicit clarification through the dialogue may serve as a marker or

FIGURE 9.3A

Con Air. Inmates dance and sing as they escape. [01:29:30]

FIGURE 9.3B

Dramatic irony in Con Air.

Categorization

129

indicator of ironic intent. In Con Air, several notoriously violent ex-convicts take control of a plane and escape. Their celebration involves dancing and singing to “Sweet Home Alabama”, but most of the characters are not aware of the tragic (rather than joyful) connotations of the music. In a somewhat subdued voice, the more intellectual of the criminals, a serial killer, quietly utters, “Define irony: Bunch of idiots dancing on a plane to a song made famous by a band that died in a plane crash.” The plane ultimately crashes, and everyone but the serial killer dies in the accident. Irony thus plays on a disjunction between the characters’ and the audience’s points of view, with a flagrant incongruity between the song’s connotations and the convicts’ state of happiness. The reversal, however, takes place in the audience’s realization of the convicts’ likely fate. As Figure 9.3B shows, this is an example of ‘dramatic irony’ (often called ‘tragic irony’), where the audience knows what the character has yet to find out. Stereotyped characters commonly share their looks, general behavior, and even musical taste. White Chicks brings stereotypes to the foreground only to deconstruct them for humorous effect. As a group of young white girls rides in a convertible through an upperclass neighborhood, the radio commentator announces, “And now, the number one most requested song at WQQR”, and Vanessa Carlton’s “A Thousand Miles” begins to play. The girls react with great excitement, “This is our jam!”, and start singing along. In a subsequent scene, an undercover black cop disguised as a white girl, Tiffany Wilson, joins a masculine black bodybuilder, Latrell Spencer, for an after-party ride. As ‘she’ recognizes Latrell’s intimate intentions, she attempts to establish some distance by playing “A Thousand Miles” on the car stereo, a song that would turn him off. To her surprise, Latrell exclaims, “I love this song!” He starts singing along, shaking vigorously during the orchestral riffs and impersonating the lyrics. By exposing his musical taste, he finds his way into a radically different tradition. The incongruity between Tiffany’s (and our) expectations regarding Latrell’s musical taste and his (arguably exaggerated) joyful reaction to the song, combined with his hyperbolic enactment of the lyrics, causes a reversal of character type. Because this example embeds a fundamental incongruity that reveals a state of affairs different from expected, it belongs to the structural irony category, often called situational irony. (See Figure 9.4B.) Superstar portrays the school life of Mary Katherine Gallagher, a rather graceless and mildly hyperactive uniform-wearing Catholic schoolgirl with dreams of superstardom. During a Roman Catholic service set to Schubert’s peaceful “Ave Maria”, as Mary whimpers about her unsympathetic reality, her overenthusiastic friend, Helen, cues a daydream sequence with the line: “That’s it. You are feeling sad, so you know what it’s time for? Supermodel documentary hour!” Both girls leap into a daydream sequence of ‘superstardom’ as supermodels in a photo-shoot session set to Imani Coppola’s “I’m a Tree”. This dazzling moment frees them from the characteristic awkwardness that brands them as social outcasts in their real lives. Yet, they also adopt hyperbolic and grotesque mannerisms that caricature supermodels, such as the distorted ‘French’ accent Helen enacts when recalling, “I was just walking down the street one day and a man come up to me and he said, ‘Do you like to be a supermodel?’, and I said oui, and the next day I’m in New York, on the cover of Vogue.” This dreamlike sequence is filled with cinematic anomalies—rapid camera movement, constant white-screen bursts simulating a camera’s flash. The transition from their dreamlike state to reality is affected by a visual element—a priest unexpectedly appears within the

130

Categorization

FIGURE 9.4A

White Chicks. Latrell’s reversal of character type. [01:02:50]

FIGURE 9.4B

Structural irony in White Chicks.

dream sequence—prompting Mary and Helen to return to their nervous stillness while Mary murmurs her characteristic ‘Sorry, sorry.’ These cinematic cues serve as indexical signs that unmask the director’s intention: to target Mary and Helen via a hyperbolic and grotesque statement of the opposite. As Figure 9.5B shows, the blatant ‘blame-by-praise’ and presence of a target of ridicule or critique in this example suggests it is an instance of Socratic irony; and because the target is a character in the film, rather than social norms or the film genre itself, this example may be further read as an instance of sarcasm. The drama Precious revolves around a young, black, overweight, sexually and emotionally abused girl. Daydreaming about an alternative life and a different identity are her only means of coping with the grim reality surrounding her. Leading to one of the many dreamlike sequences, we see Precious walking through her neighborhood past a group of harassing bullies hearing some indistinct R&B music. One guy pushes Precious to the ground. As she falls unconscious, the film cuts to a dreamlike sequence: she wears a glamorous dress

Categorization

FIGURE 9.5A

Superstar. Mary and Helen daydream of superstardom. [00:19:45]

FIGURE 9.5B

Socratic irony and sarcasm in Superstar.

131

and dances on a stage with her imaginary light-skinned boyfriend to Queen Latifah’s “Come Into My House”. As her (imaginary) boyfriend gets behind her and licks her ear, the film cuts to reality, and we see a dog licking her ear while she is on the ground regaining consciousness. The cinematic anomalies in the scene—the split-screen montage typical of music videos, a shift to vivid colors, slow-motion and stop-motion editing—serve as indexical signs that unmask the director’s intention: targeting Precious via a hyperbolic statement that plays on a transmuted expression of reality, one of excess and glamour. This cinematic device calls attention to Precious’s bleak life, which sadly lacks the glamor she craves. Such hyperbolic and grotesque admiration in the form of ‘blame-by-praise’ is widespread in discursive irony, tacitly implying the word ‘not’ as a semantic modifier. The presence of a target of critique or ridicule suggests this is an instance of Socratic irony; and because the target is a character in the film rather than the film genre itself, we may read this moment as an instance of sarcasm. (See Figure 9.6B.) However, within the context of the entire film, this moment of sarcasm

132

Categorization

FIGURE 9.6A

Precious. Precious’s alternative reality. [00:20:15]

FIGURE 9.6B

Socratic irony, sarcasm, and satire in Precious.

serves to project a broader critical judgment on the social dynamics that caused Precious’s reality, thus resulting in satire.10 Guess Who is a romantic comedy that touches upon racial issues. Simon, a young white man, plans to marry the daughter of Percy, a protective African-American dad. After a somewhat heated discussion, Percy decides Simon should stay in a hotel. Unfortunately, the hotel is fully booked, so they drive back to Percy’s home. As they get in the car, the ending of the song “Ebony and Ivory”, performed by Paul McCartney and Stevie Wonder, is playing on the radio, and the lyrics, “Ebony, ivory, living in perfect harmony”, make both Simon and Percy uneasy. The conflict between the associations brought about by the lyrics of the song (itself a metaphor about the white and black keys of the piano arranged in perfect harmony) and the events in the narrative (suggesting a fundamental lack of harmony between Simon and Percy) elicit an ironic reversal of the meaning of the song. Just as in the last two examples, this ironic reversal is characterized by a ‘blame-by-praise’, suggesting this too is an example of Socratic irony. (See Figure 9.7B.) In this instance, however, the critique is not directed at a character in the film, but at

Categorization

FIGURE 9.7A

Guess Who. Percy and Simon drive back home. [00:29:45]

FIGURE 9.7B

Socratic irony and satire in Guess Who.

133

the macro-social dynamics and a contextual situation, further suggesting that this is an instance of satire. Using this rhetorical device, the film shifts its narrative agency from the characters to a tacit narrator, one that allows for a shift in perspective and a broader critical appraisal of the social dynamics unfolding in the scene. A comparable example, but from a very different genre, appears in Michael Moore’s Bowling for Columbine. This documentary film is a multi-layered examination of the Columbine tragedy and a stark critique of gun ownership in the United States. In a lengthy montage, accompanied by Louis Armstrong’s “What a Wonderful World”, the film recounts the United States’ involvement in foreign and military policies that may have led to 9/11. The coupling of violent visuals with a song whose positive lyrics point to a trouble-free world illustrates Moore’s signature sense of Socratic irony employed for satirical purposes to advance a critical appraisal of the social norms and the political climate that led to a collective state of fear and paranoia. (See Figure 9.8B)

134

Categorization

FIGURE 9.8A

Bowling for Columbine. Montage accompanied by “What a Wonderful World”. [00:24:30]

FIGURE 9.8B

Socratic irony and satire in Bowling for Columbine.

The comedy Repossessed alludes to an earlier horror film, The Exorcist, by presenting many inter-textual relations—setting, iconography, character types, and even a common actress in an identical role (Linda Blair). In a scene, Father Jedediah plans to rid Nancy of the devil that possesses her. He quietly enters Nancy’s bedroom and slowly approaches her as she is tied to the bed. As he gets closer and closer to Nancy, the music grows louder and louder; but he suddenly presses a button on a handheld device and causes the music to stop. Allowing a character to have diegetic control over the non-diegetic music is an anomaly that functions as a marker, a punchline that deconstructs the cinematic illusion, triggering a reversal of effect. As we saw in Chapter 3, an increase in loudness usually triggers tension in the audience and points to an imminent dangerous event. Here, however, the unusual transgression of sound design boundaries, particularly at a sonically climactic moment, releases the tension and evokes laughter. This example aligns with parody, combining hyperbole and the grotesque to critically appraise and caricaturize a film (The

Categorization

FIGURE 9.9A

Repossessed. Father Luke attempts to exorcise Nancy. [00:22:20]

FIGURE 9.9B

Parody in Repossessed.

135

Exorcist) and the horror genre writ large. (See Figure 9.9B.) Although there is a reversal of effect (laughter rather than suspense), there is no Socratic blame-by-praise structure or significant incongruity. Hence, this example should not be read as an instance of irony, but as an instance of parody. In the comedy Airplane!, a reconfiguration of the symbolic system generates a vacillation between similarity and difference in relation to other films. It opens with a shot of clouds seen from slightly above, almost resembling a turbulent sea. Jaws’ menacing theme emerges as the plane’s ‘fin’ breaks through the clouds. As the music builds in intensity, the plane’s fin gets closer. Ultimately, as the entire plane cuts through the clouds in a sharp ascent, a shocking dissonant chord in the music functions as an anomaly that punctuates the scene with a musical gag-line that intensifies the humorous effect, landing a biting example of parody. (See Figure 9.10B.)

136

Categorization

FIGURE 9.10A

Airplane! A plane’s ‘fin’ emerges from the clouds. [00:00:05]

FIGURE 9.10B

Parody in Airplane!

In the dark comedy Ted, there are copious references to 1980s films. In a scene filled with intrepid allusions to Indiana Jones’ musical theme, just like Indy grabs his hat while making his grand escape, Ted grabs his ‘cloth’ ear before his adventurous flee down the stairs. Although intertextuality is the most salient feature here, there is no reversal of effect, no anomaly, and no element of the grotesque; therefore, this moment in Ted engenders a sense of homage to Indiana Jones and thus would be best categorized as quotation rather than parody. (See Figure 9.11B.) Face Off is an action/crime/sci-fi film in which an FBI agent undergoes facial transplant surgery to assume the identity of a terrorist. In a scene, a child witnesses a violent event. Attempting to distance the child from this violence, the FBI agent places a set of headphones over the child’s ears playing a gentle version of “Over the Rainbow”. Colored by the song, the violence turns into exquisitely choreographed dance-like movements. The incongruity

Categorization

FIGURE 9.11A

Ted. Ted channels his inner Indiana Jones. [01:21:00]

FIGURE 9.11B

Quotation in Ted.

137

between the music and the suggestion of violence in the visuals and narrative does not trigger an ironic reading of the scene, but instead, the scene portrays the child’s perspective. As Figure 9.12B shows, this is an example of a paradox, which entails a contradiction that makes sense and does not require resolution. Meet the Parents is a comedy about Greg meeting his prospective in-laws. Jack, the disapproving father-in-law, has one precious possession: Jinx, a white-tailed, HimalayanPersian cat. During Greg’s weekend with the in-laws, Jinx is nowhere to be found. To gain Jack’s respect, Greg disguises a newly acquired cat by painting its tail white and pretends to have found Jinx. As Greg arrives with the disguised cat, the majestic and heroic flair of the music (via a ‘heroic’ musical topic) corresponds with the visuals of Greg as determined and strong-minded (via the characteristic slow-motion and a half-smile). The film thus presents us with contradictory depictions of Greg: the plot presents him as hopeless and illfated, while the hyperbolic music and visuals portray him as a hero. Here, however, the

138

Categorization

FIGURE 9.12A

Face Off. The child is caught in the midst of the violence. [01:34:10]

FIGURE 9.12B

Paradox in Face Off.

inter-textual reference to the epic genre does not establish parody, because the genre is not critically appraised. Additionally, no second-order interpretation emerges—no reversal of effect, critical appraisal, shift of focal point, or homage. Moreover, although Greg’s fate is not a blissful one, no narrative elements in the film point to or suggest that outcome. Therefore, as captured in Figure 9.13B, this example does not fit irony, satire, sarcasm, or parody; instead, and although music is incapable of lying, this example suggests that the music helps the character lie by staging a false heroic persona. Terminator 2: Judgment Day builds on the character stereotypes developed in The Terminator; but in a role reversal, a reprogrammed T-800 (the one we identify as ‘The Terminator’) arrives to protect John Connor from a T-1000, a more advanced, shape-shifting android assassin. In the scene from this film with which I open the chapter, the Terminator seizes a rugged-looking biker’s clothes, steps outside the diner, and even grabs the biker’s Harley.

Categorization

FIGURE 9.13A

Meet the Parents. Greg arrives with a lookalike cat. [01:13:10]

FIGURE 9.13B

Lie in Meet the Parents.

139

“Bad to the Bone” rumbles loudly on the soundtrack. The diner’s owner comes out with a 10-gauge Winchester lever-action shotgun, fires a round, and threatens, “I can’t let you take the man’s wheels, son. Now get off, or I’ll put you down.” Expressionless, the Terminator eases the bike onto its kickstand and strides toward the guy. Staring impassively, the Terminator snatches the shotgun—the guy gulps, thinking he is doomed. The Terminator calmly reaches toward the man’s shirt pocket, grabs his sunglasses, and puts them on. Now looking the part, the Terminator steps back on the Harley and roars off. Here, the non-diegetic song helps develop the character’s tough-guy facade, constructing a lie on two levels—that the Terminator is human and a “bad” guy. (See Figure 9.14B.) Some combinations of first- and second-order interpretations do not reflect any of the tropes discussed. In The Grand Budapest Hotel, ludicrous characters engage in un-naturalistic conversations that match their preposterous behavior. M. Gustave clasps Madame D’s

140

Categorization

FIGURE 9.14A

Terminator 2: Judgment Day. The Terminator arrives. [00:09:15]

FIGURE 9.14B

Lie in Terminator 2.

hands while comforting her, “You’ve nothing to fear. You’re always anxious before you travel. I admit you appear to be suffering a more acute attack on this occasion, but truly and honestly [abrupt stop]”—he is suddenly taken aback by Madame D’s “diabolical” nail varnish. At that very moment, the non-diegetic music also abruptly stops, and M. Gustave’s tone shifts from affectionate and reassuring to critical and disapproving. Throughout the scene, the film becomes Gustave’s accomplice, a supporting agent or unseen narrator that supplies the appropriate non-diegetic music to bolster the romantic narrative Gustave is trying to foist upon Madame D. The halt in the music, however, draws attention to the dialogue, and farcically complicates the presence of the non-diegetic music as influenced by the dialogue itself, foregrounding cinematic conventions only to deconstruct them for humorous effects. Unfortunately, the categories here discussed do not fit a humorous trope that entails incongruity based on a cinematic anomaly—a sudden stop in the non-diegetic music that brings with it an equally sudden change of valence in the dialogue. (See Figure 9.15B.)

Categorization

141

FIGURE 9.15A

The Grand Budapest Hotel. M. Gustave and Madame D. talk before her departure. [00:10:10]

FIGURE 9.15B

Indeterminate rhetorical trope in The Grand Budapest Hotel.

Fluid Categories

Identifying the attributes (first- and second-order interpretations) that inform our categorization, as presented in Figure 9.16, allows us to distill the difference between different shades of irony and between irony and related tropes. Related to first-order interpretations: (1) incongruity is a necessary attribute in irony, but the location of such incongruity may suggest a non-ironic trope; (2) intertextuality rules out ironic meaning; and (3) hyperbole, anomaly, and the grotesque may be present in ironic as well as non-ironic tropes. Related to second-order interpretations: (1) reversal of effect is pervasive in irony and parody, but is not as common in tropes distant from irony; (2) critical appraisal requires further examination to identify whether sarcasm, satire, or parody are at play;11 (3) a focal shift is not common

142

Categorization

FIGURE 9.16

Exemplars, attributes, and categorization.

in irony, but it appears more often in distant tropes such as paradox; (4) homage emerges as a distinct qualifier for quotation, (5) the absence of second-order interpretations may indicate a lie; and (6) the presence of first- and second-order interpretations notwithstanding, additional markers may suggest an undefined trope, one for which we may not yet have a fitting label. Since the framework for categorizing irony and related tropes I present in this chapter is based on graded similarity to a prototype (or, often, to an exemplar) rather than on explicitly specifying the boundaries between categories or abstracting the necessary and sufficient conditions for category membership, knowledge of categories demands knowledge of near-prototypical exemplars. Confined to the examples provided in this chapter, knowledge of the category parody, for instance, comprises knowledge of near-prototypical examples such as those from Airplane! and Repossessed. The prototype model thus offers a suitable approach to distilling semantic categories featuring fluid boundaries and typicality-based category membership.12 Nonetheless, the members of any semantic category are rarely a perfect prototype, as these may deviate in the degree of membership and even contain cues suggesting closely related or even distant categories. It would therefore be theoretically advantageous to conceive of irony and other tropes within a multidimensional continuum, a space that situates categories relative to each other and allows for inbetween-trope graded typicality or regions of overlap where non-prototypical exemplars could be situated. Figure 9.17 renders such multidimensional space, wherein distances between categories are relative. Over time, new categories (along with their corresponding prototypes) may emerge from within interpretive communities—for instance, to label the trope in The Grand

Categorization

FIGURE 9.17

143

Multidimensional mapping of categories.

Budapest Hotel—forcing existing exemplars to shift in their position closer or away from a prototype. This multidimensional mapping may allow for tracing diachronic fluctuations in our categorization, continuously adjusting the location of various tropes to represent the complexity of our experience and to better approximate new exemplars. Coda

Music operates at an almost subliminal level during a film, conveying complex messages that support, highlight, or complement other cinematic domains. Occasionally, however, the music and other cinematic domains are at odds, supplying incongruous information that violates our expectations, triggering a momentary discomfort we resolve by physiologically resorting to laughter and rationally reducing the event to an instance of irony. This reflects that our everyday tendencies for basic-level categorization permeate our cinematic experiences.13 In this chapter, we probed such tendencies by cross-examining the boundaries between musically induced irony and related tropes, and proposed that new cinematic instances prompt us to recalibrate our basic-level categorization and even to contrive new categories. Terminator 2: Judgment Day also probes our innate tendency to categorization: in its soundtrack and cinematography, the film blurs generic boundaries between sci-fi, tech-noir, action, comedy, adventure, horror, fantasy, mystery, and thriller, becoming a blueprint for many future films within these genres; and, in its themes, the film portends a unique vision of a world, one that blurs the boundaries between humans and machines, protagonists and antagonists, emotion and reason, familiar and unfamiliar, memories and dreams, fate and future, being and becoming.

144

Categorization

Notes 1. When acknowledging that ‘conflicting meanings’ is a common trait in all instances of irony, it may be tempting to adopt a ‘family resemblance’ model—where category members share specific features, yet no single feature is common to all instances. 2. Although affordances and function have commonly been included as attributes, including interpretations as attributes renders a more robust model for categorizing irony. The attributes that inform our categorizations are distributed across a wide network of perspectives, including not only how items “look, but also they sound, smell and feel, how to operate them, and emotions they arouse” (Barsalou et al., 2003, p. 88). In addition, because attributes are in themselves complex concepts with both internal and external structures, the ability to establish inter-attribute relations allows us to identify attributes that may co-occur. For example, having wings, flying, and building nests in trees are closely interrelated, as having wings affords a bird flying and hence building a nest in a tree. For an in-depth discussion of the notion of categorization, see Appendix VIII in this volume. 3. Cinematic anomalies often act as indexical markers that help unmask the director’s intention, replacing typical communicative signals such as rolling eyes, intonation, or facial gestures. 4. In discussing irony and related tropes, Hutcheon (1985) suggests that parody is not located in the text, but in the readers’ minds. Elleström (2002) further distills the objective and subjective components in arriving at an interpretation, noting that “the material of irony is found in the text, but it is formed by the reader” (p. 49). 5. This list of attributes arose organically from examining the film examples included in this chapter; additional examples will no doubt contribute to expanding this preliminary list of attributes. 6. In this process of categorization, not all attributes enjoy equal weight. For instance, when constructing interpretations of parody or satire, viewers must selectively direct their attention toward the presence of intertextuality and the values of critical appraisal. In contrast, to decide between parody and homage, viewers must selectively direct their attention toward the presence of reversal of effect and the values of critical appraisal. In addition, the weights given to both the attributes and the values within them will contribute to identifying a prototype that serves as representative both within the ‘vertical’ and ‘horizontal’ dimensions of a category. The horizontality of categorization is manifested in the various tropes related to irony, all arising as variations of the incidence of the one essential element: a conflict of messages. 7. Some readers may argue that these labels fail to reflect the true nature of the proposed prototypes. I return to this issue toward the end of the chapter. Additionally, as an audience member who has worked with these examples for extended times, I commonly fall victim to learning, priming, and plain habituation. 8. Hutcheon (1985) notes that parody is “related to burlesque, travesty, pastiche, plagiarism, quotation, and allusion, but remains distinct from them. It shares with them a restriction of focus: its interpretation is always of another discursive text. The ethos of that act of repetition can vary, but its ‘target’ is always [intra-opus] in this sense”, and in turn, satire is “[extra-opus] (social, moral) in its ameliorative aim” (p. 43). Dovetailing on Hutcheon’s ideas, Cobo (2003) suggests that “the difference between parody and satire resides on the distinct nature of their respective targets . . . the parodist target is always another work of art or more generally, another form of coded discourse . . . satire, on the other hand, is both moral and social in its focus and ameliorative in its intention” (p. 61). 9. Cherlin (2017) delves into the numerous forms of musical irony found in classical music, examining works by a range of composers including Mozart and Mahler. Kostka (2016) discusses the concept of parody in modern and postmodern art, particularly as it stems from blending stylistic traditions. 10. Socratic irony stems from a particular rhetorical pattern in Socratic dialogues, in which Socrates pretends to need information and professes admiration for the wisdom of his companion. For instance, saying “Congratulations! You’re so smart . . .” to someone that just received a failing grade would be understood as ‘not so smart’. 11. Everett (2004) draws attention to this fact and defines “parody in musical discourse as a composer’s appropriation of pre-existing music with intent to highlight it in a significant way . . . The analyst then determines whether the accompanying ethos is deferential (neutral), ridiculing (satirical), or contradictory (ironic) based on how the new context transforms and/or subverts the topical/expressive meaning of the borrowed element” (par. 1).

Categorization

145

12. As Rosch (1978) mentions, the prototype model seeks to “achieve separateness and clarity of actually continuous categories [by] conceiving of each category in terms of its clear cases rather than its boundaries” (p. 259). 13. In contrast to natural categories, irony presents a unique case of basic-level categorization that remains largely abstract and disembodied at subordinate levels—e.g., romantic irony, dramatic irony, or cosmic irony, all retain an equal degree of abstraction compared to the superordinate level.

10

INTERPRETIVE TRANSFORMATIONS

In The Red Violin, the music invites us on a transformative journey, one that traverses over three centuries and several continents. The extraordinary, bewitching, desired red violin is in line for auction. As the bidding unfolds, a series of flashbacks brings us along with the violin from its creation in seventeenth-century Italy to its ever-changing milieu as it falls in the hands of peculiar individuals—a child prodigy in an eighteenth-century Austrian monastery, a Romani female improviser, a mischievous soloist in nineteenth-century Oxford, a violin restaurateur in China during the Cultural Revolution, and an appraiser of antique instruments in present-day Montreal. Sweeping transformations of the ‘red violin’ leitmotif guide us through the transient owners’ sociocultural environments while illuminating their inharmonious relationships with the instrument, with life, with music. In this chapter, we dovetail on the themes from previous chapters and fine-tune our understanding of leitmotifs, topics, and musical associations, now attending to their transformations and the impact of such transformations on our interpretations of the film’s dramatic trajectory. First, we explore instances where leitmotif variations indicate an analogous affective or associative realignment in the storyworld. Then, we investigate musical troping, a technique with remarkable expressive potential entailing a transmutation of unrelated topics. Last, we examine how a reconfiguration of a song, or the context surrounding a song, prompts us to reframe our interpretations of the music’s place within the narrative. Leitmotif Transformation

Leitmotifs are fluid, changeable constructs whose transformations help outline and explain narrative or character developments. Once established, leitmotifs cue our attention to specific elements in the storyworld, and, when transformed, they suggest a new affective state or supply a new semantic layer that modulates our perception and characterization of those elements.1 Films draw primarily on two types of leitmotif transformations to influence our interpretation of the narrative: those grounded in embodied mechanisms and those driven by semiotic associations.

DOI: 10.4324/9780429504457-11

Interpretive Transformations

147

Transformations via Embodied Mechanisms

The Adventures of Robin Hood (1938) offers a landmark example of leitmotif transformation via embodied schemas.2 The film unfolds in the twelfth century, when the mighty Normans ruled England with an iron fist. In King Richard’s absence, his brother, Prince John, and his henchman, Sir Guy of Gisbournethe, seize the opportunity to subjugate the Saxon peasants, taxing them beyond reasonable limits. Robin Hood, a Saxon lord himself loyal to King Richard, assembles an army, his band of Merry Men, to bring justice to his people and overthrow Prince John. Because the film sets the story in the distant past, and because the film was produced relatively early in movie soundtrack history, the filmmakers had only a limited repertoire of film music topics to draw from. Therefore, leitmotif transformations in this film chiefly draw on embodied mechanisms, with only sporadic excursions onto topical adaptations. The ‘Merry Men’ proto-leitmotif appears over the main title credits. (See Figure 10.1A.) This first iteration, rendered in duple meter, with boisterous dynamics, downbeat accents, and muscular brass and strings, infuses the music with heroic and spirited undertones characteristic of a ‘heroic march’ topic.3 Robin Hood must assemble a band of men to challenge Prince John. Transformations of the ‘Merry Men’ leitmotif will help characterize the new members’ personalities and, at times, their physical traits. The first candidate to enlist, Much the Miller’s Son, conducts himself according to ethically questionable norms. For instance, without thinking of the consequences, Much kills a deer—a grave offense within the royal forest laws. As the Norman overlords are about to bring justice to Much, Robin Hood intervenes and rescues him

FIGURE 10.1A

‘Merry Men’ leitmotif. (Music by Erich W. Korngold.)

FIGURE 10.1B

The Adventures of Robin Hood. Main title sequence.

148

Interpretive Transformations

FIGURE 10.2

The Adventures of Robin Hood. Much the Miller’s Son joins the band. [00:04: 30]

from the harsh punishment. Much, in need of moral guidance, eagerly requests to join the Merry Men, swearing, “From this day on, I’ll follow only you . . . Take me as your servant.” On his horse, looking down on the short fellow, Robin takes a few moments to respond, hesitatingly accepting him into the group. In the soundtrack, a disjointed, hesitant version of the ‘Merry Men’ leitmotif denotes the addition of a character to the band whom Robin Hood did not intend to include—short fragments of the leitmotif elicit in us a feeling of incoherence and hesitation, while the wandering tonal center affords us only uncertain and unconvincing grounds. Moments later, Robin encounters Little John, a giant man in worn clothes, carrying a quarterstaff. Each man stands on the opposite ends of a bridge, measuring themselves by one another. Realizing ‘Little’ John’s intimidating appearance, Robin immediately contemplates including him in his army, but first, he puts Little John to the test. The two men battle with quarterstaffs on the bridge—striking, ducking, reeling, but neither giving ground. Robin loses his footing and falls to the water. Little John roars with laughter and thrusts his quarterstaff to Robin to help him to the bank on the side of the bridge. Both men laugh and introduce themselves. In the soundtrack, the ‘Merry Men’ leitmotif emerges with lighthearted, comedic undertones—a duet of bassoons in their high registers—anticipating Robin’s invitation to Little John to join the band by approvingly remarking, “If you can hold a breach like you held the bridge, you’re one of us, and welcome.” As Friar Tuck joins the band, the ‘Merry Men’ leitmotif sounds in a solo bassoon playing in its high register and accompanied by pizzicato strings. Here, the instrument not only feeds humorous undertones into the musical texture—as Robin promises to the large fellow “a venison pastie and the biggest you ever ate . . . boar’s head, beef, casks of ale”—but also onomatopoeically mimics the distinctly high nasal tone of Friar Tuck.

FIGURE 10.3

The Adventures of Robin Hood. Little John joins the band. [00:23:00]

Interpretive Transformations

FIGURE 10.4

149

The Adventures of Robin Hood. Friar Tuck joins the band. [00:32:00]

As additional members join the band, additional instruments join the leitmotif (woodwinds, brass, strings, percussion), consolidating the orchestral ensemble, representing Robin Hood’s strengthening forces. From this point in the film, the ‘Merry Men’ leitmotif underscores scenes where they take on the establishment. Now complete, the band of Merry Men prepares to ambush Sir Guy’s Party. Eagerly, the men climb through the many branches into the treetops and prepare to strike. In the soundtrack, the ‘Merry Men’ leitmotif resonates fervently, with many contrapuntal voices moving in all directions yet supporting the collective texture with an orchestral tour de force. After the successful ambush, the men celebrate with a banquet at Robin’s camp—long tables, two whole steers and several pigs roasting over fire pits, outlaws celebrating and deservedly enjoying themselves dancing. In the soundtrack, the entire orchestra gleefully joins in, setting the ‘Merry Men’ leitmotif in a joyous triple meter, turning it into a delicious waltz that distances us from the leitmotif’s march-like associations presented so far in the film, affording us an engagement with the scene’s playful, leisurely, festive atmosphere.

FIGURE 10.5

The Adventures of Robin Hood. The band prepares to ambush Sir Guy’s Party. [00:33:30]

FIGURE 10.6

The Adventures of Robin Hood. The Merry Men celebrate their victory. [00:39:50]

150

Interpretive Transformations

FIGURE 10.7

The Adventures of Robin Hood. Disguised Merry Men marching toward the Nottingham Castle. [01:29:50]

FIGURE 10.8

The Adventures of Robin Hood. Ceremonial trumpets announce Prince John’s arrival. [01:31:45]

Near the film’s end, the band plots to disguise themselves as members of the Bishop of the Black Canon’s entourage to gain access to Nottingham Castle. As the men approach the gate, a camouflaged version of the ‘Merry Men’ leitmotif seeps through the soundtrack, supplying only a few structural notes, concealing its very identity. Its slow pace matches the anticipatory yet measured speed of the procession, and the tolling bells, minor mode, and dominant pedal tone (not affording us a resolution to tonic) elicit a suspenseful atmosphere that foreshadows the potentially treacherous situation ahead for the Merry Men. Robin and his men enter Nottingham Castle and sneak into the royal box. A fanfare of ceremonial trumpets announces Prince John’s arrival, as he strolls through the chamber and gallantly positions himself on the throne, unaware that the Merry Men surround him. A close listen to the soundtrack in the scene, however, reveals a most astute transformation, one that involves tinkering with the soundtrack’s diegesis via a transference—the ‘Merry Men’ leitmotif infiltrates the film’s storyworld and even the royal fanfare of ceremonial trumpets. This canny transference suggests that the band of men is poised to dominate in the clash about to unfold and even suggests a potential transformation of the prince’s forces, a realignment to the ideals of justice championed by the Merry Men. Transformations via Musical Archetypes

Musical topics are cultural archetypes, mental constructs we form upon repeated exposure to concomitant relationships between musical and extra-musical domains. Therefore, leitmotif transformations that draw on topics rest on the listeners’ enculturation to generate

Interpretive Transformations

151

meaning. In such leitmotif transformations, the intra- and extra-opus associative powers of leitmotifs and topics merge to guide us through the film’s dramatic trajectory. This subsection also draws on a single film, The Red Violin, whose soundtrack masterfully realizes a large-scale structural plan that maps the narrative’s events through topical leitmotif transformations.4 The Red Violin chronicles the journey of a mysterious instrument through a series of flashbacks, each meticulously crafted as a historical tableau that reveals the complex interrelationship between the instrument, its owners, and their cultural milieus. These flashbacks are linked by a fortune teller reading the tarot cards to Anna, the violin maker’s wife. The cards, however, do not predict Anna’s fate, but the violin’s. Along this journey, the instrument gets buried, stolen, shot, and nearly burned, but when played, it resonates with its current owner’s unique temperament to produce sublime sounds. The film begins with a present-day violin auction in Montreal but soon flashes back to the instrument’s creation in seventeenth-century Cremona. In this first flashback, Anna hums a lullaby-like tune to her child in the womb, a soothing yet nostalgic tune beginning in the Dorian mode and gently shifting to the relative Aeolian mode. Sadly, they both die during the delivery. To keep their memory alive, Anna’s husband, master luthier Bussotti, varnishes the instrument with their blood. In the soundtrack, Anna’s humming fades away and dovetails into the same tune performed on a violin—the ‘red violin’ leitmotif is born. The tarot cards predict a long journey with blissful and tragic moments. In the second flashback, the violin embarks on this journey, drifting toward an Austrian orphanage, falling in the hands of Kaspar, an orphan who attempts to learn to play the instrument. During the day, he practices tirelessly; during the night, while asleep, he holds his newly found treasure close. In this first transformation, the leitmotif adopts a ‘Baroque style’ topic, characterized by clear diatonic harmonic progressions and continuous rhythmic figurations. Despite Kaspar’s attempt to control the musical time by rehearsing to the beat of a metronome, the beat of his own heart betrays him, and he collapses while auditioning before the prince. Kaspar’s untimely death, therefore, stems from his obsessive drive (but sheer inability) to control the

FIGURE 10.9A

The ‘Red Violin’ leitmotif, as hummed by Anna. (Music by John Corigliano.)

FIGURE 10.9B

The Red Violin. The Red Violin appears for the first time. [00:06:30]

152

Interpretive Transformations

FIGURE 10.10A

‘Baroque style’ topical transformation of the ‘Red Violin’ leitmotif. (Music by John Corigliano.)

FIGURE 10.10B

The Red Violin. Kaspar’s relentless practice. [00:39:30]

instrument, the music, or even himself. Kaspar is buried right outside the orphanage, holding tightly to the violin. In the third flashback, a few years later, a Romani group expropriates the violin from Kaspar’s grave. They travel through Europe, wandering extemporaneously through many countries. In the hands of a female Romani improviser, the leitmotif re-emerges with a free spirit, infused with ad libitum gestures and exotic modes archetypical of a ‘Romani’ topic.

FIGURE 10.11A

‘Romani’ topical transformation of the ‘Red Violin’ leitmotif. (Music by John Corigliano.)

FIGURE 10.11B

The Red Violin. Romani woman improvises on the instrument. [00:51:20]

Interpretive Transformations

153

FIGURE 10.12A

‘Paganini-esque’ topical transformation of the ‘Red Violin’ leitmotif. (Music by John Corigliano.)

FIGURE 10.12B

The Red Violin. Lord Frederick Pope’s emotionally charged performance. [00:56:15]

The Romani group travels through Oxford, through the lands of Lord Frederick Pope, a nineteenth-century violin virtuoso, who takes possession of the instrument. In this fourth flashback, the leitmotif flirts with a ‘Paganini-esque’ topic to exploit Pope’s devilishly virtuosic technical capacities—ricochet bowing, multiple stopping at dazzling speeds, extended octave playing—and to allow him to exude his intense emotional and artistic expression through the music. He becomes aesthetically transfixed by the instrument and begins to include it not only in his performances, but even in his lovemaking. This makes his mistress jealous. As she finds Pope in a ménage à trois with another mistress and the violin, she fires a bullet at the instrument. One of Pope’s servants travels to Shanghai and pawns the broken instrument to an antiquarian, who repairs the damage. Although physically restored, the violin cannot resist the collective—in China, during Mao’s Cultural Revolution, the violin is portrayed as a symbol of Western decadence. Within this tableau, the oppressive sociocultural ideology entirely silences, rather than transforms, the ‘red violin’ leitmotif.

FIGURE 10.13

The Red Violin. The instrument amidst China’s Cultural Revolution. [01:16:30]

154

Interpretive Transformations

FIGURE 10.14

The Red Violin. The instrument realizes its destiny. [02:03:15]

Upon the Chinese antiquarian’s death, the government inherits the instrument and soon ships it to Montreal for auction. In present-day Montreal, Charles Morritz, an expert violin appraiser and a luthier himself, notices the much-coveted ‘red violin’, the last of legendary luthier Bussotti. Several eager buyers obsessively compete for the instrument, but merely as a trophy, rather than for its capacity to produce sublime music. Morritz, although unable to afford the multimillion-dollar instrument, is determined to procure the violin for his young daughter. In a brief flashback to the fortune teller, the final tarot card, Death, appears upsidedown, not predicting death but rebirth. Morritz exchanges a counterfeit for the real red violin during the auction. As he secures the red violin and flees the scene, he calls home to speak with his daughter and reveals, “Honey, I’m coming home, and I am bringing you something very special.” The original leitmotif, unaltered, which the audience has not heard since the film’s first moments, returns to the soundtrack, suggesting the instrument will finally realize its destiny, reaching the hands of a luthier’s child. Musical Troping

Musical troping entails fusing two distinct, unrelated topics.5 Through this fusion, the music accentuates a plot based on some dichotomy, such as two distinct characters, two different cultures, or two points of view. Whereas topics draw on conventionally established meanings, troping establishes new and localized meanings, thereby constructing metaphors that draw parallels between converging structural musical elements (source) and converging extra-musical elements (target). The examples explored in this section present a fusion of two sociocultural milieus and a corresponding musical troping in the soundtrack, here gathered under the [CONVERGENCE OF NARRATIVE ELEMENTS] IS [CONVERGENCE OF TOPICS] conceptual metaphor. Within this metaphorical process, an asymmetry between the elements entering the musical troping prompts us to map an analogous asymmetry onto the extra-musical domain. Moreover, a parallel process of signification tangential to, but enriching, the metaphorical troping is at play: identifiable pieces bring unique associations and connotations that (in some listeners may) elicit uniquely subjective readings of the scenes.6 Being There tells the story of Chauncey Gardiner, who lived in isolation, working as a gardener in a millionaire’s house. When the wealthy man dies, Chauncey must leave the house and submit himself to an unfamiliar, modern world. The film presents two contrasting elements: the inner world of Chauncey Gardiner (an old, prudent, innocent, almost

Interpretive Transformations

155

sterile human being) and the outer world (a modern and exciting, yet contaminated and corrupt environment). The soundtrack bestows Chauncey and the outer world with their own distinct musical topics: ‘classical music’ for Chauncey and ‘disco-funk’ for the outer world. As Chauncey leaves the only home he has ever known and steps into the outer world, the soundtrack fuses two unrelated topics. The musical troping in the scene blends salient markers characteristic of each topic. The music retains vital elements from a classical piece but adjusts them to conform to the ‘disco-funk’ topic: the original (symphonic) instrumentation integrates archetypical disco timbres (organ, drum set, cowbells); the melody, performed on the orchestral instruments (trumpet with sporadic orchestral tutti), adopts rhythmic nuances (anticipations and retardations) characteristic of the disco style; the harmonic progression incorporates chordal extensions (7ths, 9ths, 11ths) representative of contemporary practices; a jazzy organ breaks through the musical texture with funky riffs and counter-melodies; and, a rhythm section punctuates the metrical structure but shifts the emphasis to off-beat metrical positions and infuses a disco feel. Superimposing the resulting musical troping onto the narrative triggers a metaphorical correlation—we project the structural convergence of topics onto a parallel convergence in the narrative—placing the confluence of two widely divergent sociocultural worlds at the center of attention. A closer examination of this metaphorical correspondence, however, illustrates an asymmetrical relationship in both the musical and narrative domains—one topic enters the musical troping as dominant, subverting the other. Changes in timbre and in the sonic environment amount to the most radical transformations of the ‘classical music’ topic, now subverted and under the generic control of the ‘disco-funk’ topic—for instance, the presence of non-musical sounds (such as the city noises) and the absence of a long reverb (archetypical of concert halls) relocate the classical piece into a non-classical context. As a result, although the classical piece retains its melodic profile and other essential qualities, it surrenders its identity when entering the musical troping and becomes absorbed by the dominant ‘disco-funk’ topic. In turn, and stemming from this metaphorical mechanism, we map the musical asymmetry onto our reading of the narrative: Chauncey surrenders his identity as a gardener when leaving his old home and yields to the subduing, overpowering forces of the reckless city. In addition, a parallel meaning-making process is at play. The classical piece undergoing transformation, Richard Strauss’s “Also Sprach Zarathustra”, is a cultural trope unto itself, bringing about the “discovery of new, potentially dangerous worlds” connotations for

FIGURE 10.15

Being There. Chauncey launches himself into a new world. [00:19:00]

156

Interpretive Transformations

FIGURE 10.16

Conceptual blending in Being There.

listeners familiar with the earlier film 2001: A Space Odyssey. We project these associations onto this narrative, generating a conceptual blend that enriches our interpretation of this moment in the film. (See Figure 10.16.) Big Night is about two Italian brothers, Primo and Secondo, who emigrated from Italy and opened a restaurant in America. The restaurant is almost bankrupt, the lack of success due to the owners’ decision to keep the restaurant as authentic to the Italian tradition as possible. Here, the film establishes a dichotomy between Italian culture and American culture. The restaurant must close its doors because of a lack of customers. Secondo seeks the advice of an Italian friend who owns a prosperous restaurant called Pascal’s. The key to Pascal’s success is the fusion of the Italian and American cultures in every aspect of the restaurant, from the menu (spaghetti with meatballs) to the live performances (Italian classics with an easy-listening American twist) to its very name (an Italian name with the characteristically American apostrophe plus ‘s’). In the film, as Secondo enters this very successful restaurant, the music and the visuals portray this cultural fusion. In the soundtrack, the musical troping transforms a traditional Italian song (here representing the ‘Italian classic’ topic) by infusing archetypical markers of an ‘American easy listening’ topic: the delivery of the original Italian lyrics takes on a pronounced American accent; an electric piano accompaniment replaces the characteristic orchestral accompaniment; profuse syncopations (characteristic of American popular and jazz styles) seep through the melody’s rhythmic profile; and, the singer’s vocal delivery and performance inflections embrace an almost karaoke singing style. A metaphor thus emerges, via which we project this topical fusion onto a parallel fusion in the narrative.

Interpretive Transformations

FIGURE 10.17

157

Big Night. Secondo visits Pascal’s to learn their formula to success. [00:23:40]

A closer examination of this metaphorical correspondence also illustrates an asymmetry in both the musical and narrative domains. One topic enters the musical troping as dominant, subverting the other. Here, the resulting timbres and overall sonic environment are vital in recognizing this asymmetry—while this rendition of the traditional Italian song retains its melody and other essential qualities, non-musical sounds (of customers and servers) act as environmental markers that relocate the traditional song into a radically different context. The ‘Italian classic’ topic thus gives in to the generic control of the ‘American easy listening’ topic. Concurrently, the narrative reveals this asymmetry—for instance, by showing how Pascal’s retained essential ingredients but willingly abandoned the traditional Italian cuisine to suit the American tastes and culinary expectations, as in their spaghetti and meatballs, which Secondo views as a butchered version of an Italian classic. In the hands of the nightclub singer, the musical troping further contributes to our alignment with Secondo’s perspective by her rendering a butchered version of a classic Italian song.

FIGURE 10.18

Conceptual blending in Big Night.

158

Interpretive Transformations

In addition, by introducing a specific song, “’O Sole Mio”, the soundtrack concocts a complementary meaning-making process. The association of this song with Italian culture has been so emblematic that it replaced the Italian national anthem during the 1920 Summer Olympics when a recording of the anthem could not be found. Nevertheless, the associations brought about by “’O Sole Mio” extend beyond the anecdotal. Broadly, the lyrics of the song refer to the sun, which rises every day and brings a new beginning, which can be interpreted as a metaphor for the potential rebirth of the failing restaurant, for the potential to bring new life to their struggling business and begin anew. More specifically, we may project the lyrics “sta nfronte a te” [“right in front of you”] onto the narrative and read this moment as emphasizing both the physical location of the restaurants (one in front of the other) or as a metaphorical message to Primo and Secondo: the potential for a radical transformation, one that may salvage their ailing restaurant, lies right in front of them. However, taking that path would require them to give in to the cultural fusion embraced by Pascal’s, a compromise they are unwilling to make. (See Figure 10.18.) Saturday Night Fever explores society’s cultural, economic, and ethnic divisions. Tony is a Brooklyn native of Italian descent who aspires to cross social boundaries by entering a discodance contest partnering with Stephanie, a classical ballerina. After a night at the disco, Tony and his entourage drive Annette, another dancer, to the mythical Verrazano-Narrows Bridge, where they ritually stop to lark about.7 Drunk yet full of energy, they dance and hang from the bridge girders, daring one another to climb higher. The non-diegetic soundtrack presents a musical fusion of the ‘classical music’ and ‘disco music’ topics. Startled, Annette watches them from the car. As the music climaxes, Joey dangerously jumps, barely grabbing the rails, and Tony and Double J. leap to the rescue. Annette screams out of the car and runs to the edge of the bridge. It was a trick—the three boys, safely standing on a ledge, chant in unison, “Can you dig it? I knew that you could.” Here, the amalgam of diverse musical sounds resulting from the troping metaphorically mirrors the pluralistic dance floor.8 In this case, however, an asymmetry in the metaphorical correlation suggests an alternative reading of the scene, one in which a celebration of diversity and plurality gives way to transgression. At a structural level, the chaotic mingling of sounds hints at a breach of aural boundaries that translates into interpersonal assault—the timbres and musical gestures representing “lowbrow” disco styles (wah-wah guitar, unruly percussion, pounding electric bass) forced into the refined symphonic orchestration characteristic of “highbrow” musical styles foreshadow the sexual transgression in the gangbang scene.

FIGURE 10.19

Saturday Night Fever. Tricks of life and death on the Verrazano-Narrows Bridge. [01:10:50]

Interpretive Transformations

FIGURE 10.20

159

Conceptual blending in Saturday Night Fever.

Additionally, the programmatic aspect of a particular piece, Modest Mussorgsky’s “Night on Bald Mountain”, here included to represent the ‘classical music’ topic, enriches our reading of the scene. At a connotative level, the music borrows associations particular to this piece to depict the events in the scene. This work was inspired by Gogol mythology, specifically by a story depicting the dark witches’ Sabbath, conjuring up the devil on a desolate mountaintop. Analogously, in the harrowing nighttime scene in the film, Tony’s gang gathers on the bleak Verrazano-Narrows Bridge to play tricks of life and death, while the sacred overtones embedded in the music (synthesized bells and choir) conjure a heightened atmosphere of profanity.9 (See Figure 10.20.) Conceptual Blend Transformations

During a film, the associations elicited by a song and its lyrics impinge upon our interpretation of the narrative. In some unique cases, transformations in a song or the narrative context surrounding a song will prompt us to generate radically different conceptual blends, reconfiguring and reframing our interpretations of the music’s place within the film. In Dance of the 41, an instrumental version of a classical song plays a vital role in supporting the dramatic developments. The film chronicles the events that led to a socio-political scandal in early twentieth-century Mexico, involving the police raiding a gathering of fortytwo gay men, including Ignacio de la Torre, Mexico’s incumbent president’s son-in-law. Ignacio leads a double life, rapidly advancing in the world of politics by engaging in a loveless, lethargic marriage with the President’s daughter, Amada Díaz, while simultaneously cavorting with a young lawyer, Evaristo Rivas. The piano-only version of Schubert’s “Ständchen”, from his song cycle Schwanengesang, bookends the narrative to add radically

160

Interpretive Transformations

FIGURE 10.21A

FIGURE 10.21B

Dance of the 41. Amada attempts to master Schubert’s “Ständchen”. [00:13:20]

Conceptual blending in Dance of the 41.

different subtexts at the beginning and ending of the film. In an early scene, Amada waits for Ignacio to come home, sitting at a piano, striving to master an excerpt of Schubert’s song. As he arrives late at night, Amada continues to play. Here, the absent lyrics, “Softly my songs plead, through the night to you; down into the silent grove, beloved, come to me!”, voice Amada’s struggling plea for Ignacio’s withheld love. Despite her efforts, however, she cannot find the right tone. (See Figure 10.21B.) Toward the end of the film, Ignacio and Evaristo (and their amorous relationship) become the talk of the town, tainting Amada’s reputation in society and turning their marriage into a noxious arrangement solely intended to salvage their social status. Although smothered by the suffocating circumstances, Amada regains her agency and deliberately turns Schubert’s tender song into an inharmonious, disjointed, irritating musical pronouncement, one that voices her despair and discontent by suggesting the opposite of the song’s original meaning. Our original conceptual blend now introduces correspondences of ‘disanalogy’ between input spaces. All viewers will recognize that her deliberately distorted playing reveals how their marriage, which could have been a tender song, has become a grating, barely recognizable version of its potential idealized self. Viewers familiar with the song’s lyrics, in turn, will recognize that Amada has deliberately transformed her original soft plea into a harsh message of rejection.

Interpretive Transformations

FIGURE 10.22

161

Dance of the 41. Amada plays a transformed piano version of Schubert’s song. [01:15:20]

In the psychological drama He Loves Me, He Loves Me Not, Nat King Cole’s “L-O-V-E” brings about radically different readings of the story. In this case, it is not the music’s transformations but a perspective flip and the resulting reframing of the protagonists’ realities that prompt us to reconfigure our conceptual blends. The story follows Angélique, a talented art student madly in love with Loïc, a married man she believes will leave his pregnant wife. The film first presents Angélique’s romantic and idyllic point of view. In the bloom of love, she sends a delicate pastel-pink rose to Loïc, who smiles upon receiving it. The song “L-O-V-E” accompanies the first montage, which depicts whimsical moments of a budding romance—Loïc’s driving Angélique to his neighborhood, him playing in a park while she pencil-draws him from a distance, her sending him her art projects and him hanging them in his office. The lyrics narrate: L is for the way you look at me; O is for the only one I see; V is very, very extraordinary; E is even more than anyone that you adore can . . . Love is all that I can give to you; Love is more than just a game for two; Two in love can make it; Take my heart and please don’t break it; Love was made for me and you. We project the associations elicited by the song’s lyrics—along with its major mode, lively tempo, bouncy rhythms, and light texture—onto Angélique’s enthusiastic pursuit of love, both through her art and with Loïc. (See Figure 10.23B.)

FIGURE 10.23A

He Loves Me, He Loves Me Not. Angélique’s and Loïc’s idyllic budding romance. [00:10:30]

162

Interpretive Transformations

FIGURE 10.23B

Conceptual blending in He Loves Me, He Loves Me Not.

However, as the film unfolds, we notice that Loïc does not return her affection. Angélique’s friends note marks of probable physical abuse and warn her of a potentially unhealthy relationship. In a turning point, Loïc never arrives at the airport for a romantic getaway to Florence. Devastated, she spirals into depression. By this point, we, and Angélique’s friends, are convinced that this kindhearted, bright, impressionable young woman fell in love with a psychologically and physically abusive married man. Partway through, the film rewinds to the beginning and presents many of the same events in a new light, from Loïc’s perspective. In a heartbeat, the film’s tone shifts from romantic comedy to psychological thriller. We learn that Angélique’s infatuation sprang from a fleeting encounter with Loïc, one he barely recalls—a brief exchange of only a handful of words and his offer of a ride to a neighboring house she was looking after. As this new perspective begins to reveal her true, dark colors, he turns from suitor to victim: an obsessed, unpredictable, violent stalker (Angélique) endeavors to tear Loïc’s life apart by attempting to break up his marriage and even to murder his wife. As Loïc accidentally finds out about Angélique’s obsession, she strikes him with a brass figurine, and he falls down the stairs. The song “L-O-V-E” accompanies a second montage, which depicts Angélique being arrested, diagnosed with erotomania, and remanded to a mental institution that implements electroshock procedures, all interspersed with snapshots of Loïc unconscious at the hospital’s intensive care unit, later regaining consciousness, relearning to walk, and mending his marriage, yet hobbling around with a walker years after the incident. Primed by our first hearing of the song within a radically different context, we now notice the incongruity between the song’s lyrics and the narrative. As we subliminally attempt to resolve this incongruity, we begin to notice obsessive undertones in the song’s lyrics, ones that paint a one-sided portrait of love. Our original conceptual blend now introduces correspondences of ‘disanalogy’ between input spaces. Within this new context, the song takes on a sinister hue, creating a stark contrast to the rosy veneer it brushed over the first montage, re-casting Angélique as an unreliable narrator: while at the beginning of the film Angélique presents herself as madly in love, now we learn she is just mentally deranged.

Interpretive Transformations

FIGURE 10.24

163

He Loves Me, He Loves Me Not. Consequences of Angélique’s delusional behavior. [01:27:30]

Coda

In this chapter, we explored instances where musical or contextual transformations broaden the soundtrack’s expressive potential for meaning construction: leitmotif transformations, grounded in embodied mechanisms or driven by topical associations, suggest a new affective state or supply a new semantic layer; musical troping accentuates a dichotomy in the plot by metaphorically foregrounding a fusion of two distinct elements in the storyworld; and the associations brought about by musical or contextual transformations, particularly in relation to a song and its lyrics, prompt us to reconfigure and reframe our interpretations of the music’s place within the film. In The Red Violin, the main leitmotif becomes a latent memory, reconstructed time and time again with each transformation. As a result, the music strings together the film’s threads by transcending temporal and spatial relations, blurring physical and spiritual boundaries, and rebuilding a harmonious relationship with the past, the present, and the future. Notes 1. Just as we recognize a film’s protagonist with different haircuts or wearing different clothes, we recognize leitmotif transformations even when disguised within intricate textures. In fact, we are evolutionarily hardwired to recognize leitmotif transformations, rather than to regard these as new musical fragments. For additional insights on the ecological advantages of auditory perception, see Appendix V in this volume. 2. Winters (2007) offers a fascinating, in-depth study of Korngold’s score for The Adventures of Robin Hood. 3. Although most listeners will hear the ‘Merry Men’ leitmotif for the first time during the film, the music is itself a transformation of a waltz theme, an earlier work by the film composer himself, Erich Korngold. He composed this theme as part of an operetta, Rosen aus Florida, to fit a lighthearted scene in which a beauty contest of nations occurs. To represent Austria, Korngold composed a waltz that brings together characteristic musical figures suggesting his “homeland’s musical style” (Winters, 2007). 4. Composer John Corigliano mentions, “Because the movie was involved with the tarot, involved with classical music and a violin, and involved with many different ages of music, I felt that one had to tie everything together. If you just wrote baroque, classical, romantic, and so forth, they would be detached, since the only thing that threaded through the 300 years was this violin. One needed a thread that had to be thematic and harmonic . . . Unless you tie it together with some common thread, you will not feel the organic quality of the movie” (cited in Schelle, 1999, p. 160). That common thread is the ‘red violin’ leitmotif. 5. Hatten (1994) notes that troping occurs when “two different, formally unrelated types are brought together in the same functional location so as to spark an interpretation based on their interaction” (p. 295).

164

Interpretive Transformations

6. Within a soundtrack, there is a significant difference between incorporating an identifiable composition and simulating a generic style or topic. 7. The bridge, an American feat with an Italian name, symbolizes the realization of ambitious dreams. 8. Scholars have often pointed at the mix of disco and classical music as an expression of diversity and plurality; in the words of McLeod (2006), disco music created a “magical dance floor” that included “gays, straights, business executives, working-class heroes, whites, blacks, and Latinos, all boogying to a disco beat . . . Disco appealed to a wide range of age groups and was often the music of choice at many cross-generational gatherings such as weddings and bar mitzvahs . . . Disco united people with diverse experiences on the dance floor, participating in a shared experience of the body. In the collectivity of the dance floor, dance and dance culture manifest a politics of community, typically as a means by which individuals bind together in order to empower themselves” (p. 347). 9. For some listeners, the associations stemming from Stokowski’s arrangement of Mussorgsky’s piece as featured in Disney’s Fantasia (1940) in the scene introducing Chernabog, the evil god in Slavonic mythology, could mediate their understanding of the scene.

APPENDIX I Empathy

Through empathy, we attune to others’ actions, intentions, and emotions. In day-to-day experiences, my heart pounds, my breath quickens, I break out in a cold sweat; this physiological response elicits in me the perception of anger, causing me to gesticulate with sudden and unpredictable movements and to speak with marked shifts in intonation and faster syllabic pace; in turn, you mirror my physiological cues, channeling my speech and movements within your own body, subconsciously submitting yourself to a firsthand experience of my anger. In this appendix, I advance a process model of social empathy and flesh out its various stages, from the encoding to the decoding of affective states. Process Model of Social Empathy

Empathy is grounded in a mirroring system, a mechanism of action representation that influences our behavior. Figure I.1 outlines this mirroring mechanism: a physiological state modulates our affective states and influences our vocal/bodily gestures and postures; through a mirroring system, others simulate those vocal/bodily gestures and postures, which prompts them to experience an analogous affective state. Although the physical and psychological

FIGURE I.1

Process model of social empathy.

166

Appendix I: Empathy

attunement this model purports has not been empirically verified as a whole, much recent research in social neuropsychology and cognitive neuroscience has validated the individual mechanisms that contribute to an intersubjective resonance of physiological and affective states. Encoding Affect

We uncritically accept the idea that emotions trigger physiological states, that our mood defines how we feel physically. However, recent research in cognitive psychology indicates that this belief stems from intuitive yet misleading sensory information and suggests that the reverse chronological relationship is at play: physiological states cause emotions. William James was one of the earliest psychologists to advance this counter-intuitive proposition: Common sense says we lose our fortune, are sorry and weep; we meet a bear, are frightened and run; we are insulted by a rival, are angry and strike . . . this order of sequence is incorrect . . . the bodily manifestations must first be interposed . . . the more rational statement is that we feel sorry because we cry, angry because we strike, afraid because we tremble. (James, 1884, pp. 189–190) Keysers (2011) shows that this sequence of events is particularly evident in emotional contagion stemming from bodily mimicry; he points us to research suggesting that changes in facial expressions in others—lowering of eyebrows, clenching of jaws, raising of mouth corners—trigger in us a covert or overt mimicking of these expressions (Dimberg & Öhman, 1996), which in turn modulates of our affective states (Murphy & Zajonc, 1993).1 Analogously, Jabbi and Keysers (2008) found that observing emotions in others’ facial expressions causes a sequential, two-part response in the brain: first, we activate the cortical motor activity responsible for acting out those facial expressions; subsequently, we engage the neural substrate responsible for processing feelings and emotions (the insula). Because our affective states result from physiological changes, we may self-regulate our emotions by assuming postures or enacting gestures (Laird, 2007; Strack et al., 1988). For instance, adopting a facial expression associated with a particular emotion (anger, disgust, fear, happiness, sadness, or surprise) generates an autonomic nervous system response (heart rate variability and skin conductance) analogous to that triggered by the spontaneous facial expression (Levenson et al., 1990). Thus, “fake” and “genuine” smiles cause the same brain reaction, but to different degrees. This research supports the idea that motor activity mediates—and precedes—emotional contagion and regulation, and reminds us of the integral role of gestures and postures within a model of empathy. Gestures and Postures

There is a fundamental difference between gestures and postures in terms of their dynamic dimensions—while gestures entail movement, postures are relatively static. Both, however, are endowed with expressive significance and serve a fundamental empathic function: they reveal our physiological and psychological states via our (often subconscious) bodily cues.2

Appendix I: Empathy

167

Gallese (2017) notes that because our bodies “stag[e] subjectivity by means of a series of postures, feelings, expressions, and behaviors . . . [their] expressive content is subjectively experienced and recognized [by others]” (p. 181). Bodily gestures and postures embed cues with which to interpret their expressive purport; these cues emerge primarily from the properties of the movement itself, such as its direction, contour, speed, or force. Subtle nuances in these properties endow a meaningless movement or action with expressive power, transforming it into a meaningful gesture—a gaze becomes a stare or a flirt; a simple touch becomes a caress or a slap. Because our voices are vital instruments for expressing our affective states, we have developed a robust capacity to produce and perceive expressively nuanced vocal gestures. We imbue these vocal gestures with salient sonic features, or acoustic signatures, that correlate to particular affective states. These acoustic signatures are the very qualities that shape and transform meaningless sounds into meaningful sonic gestures—a mere exhalation becomes a lamenting sigh or a disgruntled moan. Acoustic Signatures

Physiological changes not only affect our self-perception of emotions but also modulate our vocal expressions. Respiration patterns, degrees of salivation, muscular tension, subglottal pressure, transglottal airflow, vocal fold vibration, and many other physical phenomena affect our phonation and articulation, embedding acoustic signatures within our vocalizations (Foulds-Elliott et al., 2000; Scherer et al., 2011; Sundberg et al., 2011). Our voice is thus a unique instrument that alters its sound, reflecting our emotional state. Studies on the acoustic signatures of speech and non-linguistic vocalizations identified correlations between acoustic parameters—e.g., intensity, intensity variation, spectral energy, fundamental frequency, contour, speed of syllabic onset, vibrato—and affective states—e.g., anger, fear, joy, pride, sadness, tenderness. Given the vast amount of empirical research on this topic, Scherer et al. (2017) engaged in a systematic review to distill the acoustic signatures of particular emotions. Figure I.2 synthesizes their observations for anger and sadness, illustrating a generalized phenomenon that entails the interplay between the physiological correlates of emotions and their resulting acoustic signatures.3 In addition, acoustic signatures appear to have an ecological (albeit expressive) function; these provide members of the same species with critical and “behaviorally relevant information . . . [or content] correlated directly to physical characteristics of the caller” such as sex, size, or affective state (Miller & Cohen, 2010). Larger animals, for instance, have longer vocal tracts and thus produce louder sounds with stronger fundamentals, achieving greater sound dispersion. To gain an ecological advantage, during physical conflict, “animals in an aggressive stance would try to appear larger both visually (e.g., erecting their feathers, raising their tail) and vocally, by producing lower-pitched sounds, associated with larger, stronger animals” (Eitan, 2013).4 Because a sound’s volume (or loudness) and pitch range are reliable indicators of the source’s size, our responses to these sonic parameters conform to adaptive mechanisms.5 For instance, when creating a threatening display, we produce sounds that make us appear large; but when coming across as friendly or submissive, we produce sounds that help project ourselves as small. According to this logic, and as depicted in Figure I.3, sounds representing

168

Appendix I: Empathy

FIGURE I.2

Acoustic signatures associated with anger and sadness.

FIGURE I.3

“Acoustic Ethological Model”. (After Huron et al., 2006)

large or threatening characters, objects, or situations would lie in the low-frequency range and have loud volumes, whereas sounds representing small or amiable characters, objects, or situations would lie in the high-frequency range and have soft volumes (Huron et al., 2006).6 Decoding Affect

Psychologists and philosophers have long intuited a functional mechanism of intersubjective empathy that rests on intercorporal resonance—i.e., mapping observed gestures, postures, actions, or vocalizations of others into our own bodies. James (1890) notes that “every representation of a movement awakens in some degree the actual movement which is its object” (p. 526); Lipps (1903) recalls his bodily and emotional attuning while watching an acrobat walk on a suspended wire and mentions, “in inner imitation there is no separation between the acrobat up above and me below. On the contrary, I identify myself with him.

Appendix I: Empathy

169

I feel myself in him and in his place” (as cited in Agosta, 2014, p. 61); and Merleau-Ponty (1962) notes that: the communication or comprehension of gestures come about through the reciprocity of my intentions and the gestures of others, of my gestures and intentions discernible in the conduct of other people. It is as if the other person’s intention inhabited my body and mine his. (p. 215)7 Nearly a century later, research in neuropsychology identified neural substrates that enable such intersubjective empathy: the mirror neuron system (MNS), a network of neurons that respond to both perception and action. Mirror neurons were first identified in a series of experiments on macaques;8 these neurons fired when the animal executed an action and when the animal observed an analogous action performed by the experimenter. Subsequent investigations, also in macaques, found audiovisual mirror neurons that fired when the animal performed an action and when the animal heard the related sound of that same action (Kohler et al., 2002). While the discovery of mirror neurons in macaques took place via recordings of single brain cells, non-invasive techniques have been used to confirm a homologous system in humans—the methods most widely used include functional magnetic resonance imaging (fMRI) and transcranial magnetic stimulation (TMS).9 Although these methods lack the spatial and temporal accuracy of single-cell recordings, the results gathered by these studies strongly support the existence of an action-perception link, or mirror system.10 These studies indicate that humans engage a mirror system within a wide range of conditions, subliminally activating these neural structures in response to: (1) a self-performed action or an observed/ heard action performed by someone else;11 (2) a self-experienced emotion or an observed/ heard emotion experienced by someone else; and (3) a self-experienced sensation or an observed/heard sensation experienced by someone else.12 Advancing the findings related to the human mirror system, Gallese (2010) notes that: others’ behavior is immediately meaningful because it enables a direct link to our own situated lived experience of the same behaviors, by means of processing what we perceive of others (their actions, emotions, sensations) onto the same neural assemblies presiding over our own instantiations of the same actions, emotions, and sensations. (p. 87) The human mirror system thus mediates an intersubjective process, allowing the actions, gestures, postures, sensations, and emotions of others to resonate within us, and supports a neurally grounded understanding of human empathy. While we subliminally engage the human mirror system in response to heard or seen actions, gestures, and postures, we also engage a parallel mechanism that inhibits our overtly enacting those actions, gestures, and postures (Hurley, 2008; Keysers, 2011). As Gallese (2009) explains: when the action is executed or imitated, the cortico-spinal pathway is activated, leading to the excitation of muscles and the ensuing movements. When the action is observed or

170

Appendix I: Empathy

imagined, its actual execution is inhibited . . . The development of prefrontal inhibitory mechanisms likely turns motor contagion into motor simulation. (p. 522)13 This inhibitory mechanism prevents us from automatically mimicking others’ actions, gestures, and postures, and succumbing to complete emotional contagion. On the other hand, social phenomena that engage behavioral contagion—e.g., applauding, following a crowd’s gaze, engaging in singing along—rely on the mirroring system to build a collective consciousness, setting up the conditions and environments that prompt us to relax those inhibitory mechanisms, allowing us to engage physically and emotionally with a crowd of strangers.14 Arguably, therefore, relaxing the inhibitory mechanisms in certain conditions has its roots in our evolutionary history, conditioning us to respond to a crowd’s behavior to promote ingroup bonding and ensure survival (De Waal, 2007).15 Like other submechanisms of the human MNS, behavioral contagion may be elicited both through visual and auditory stimuli, and it may reveal itself in overt actions, or it may remain covert. Moreover, group size is a strong determinant of behavioral contagion’s effect—the larger the crowd looking up, the stronger our impulse to mirror its behavior.16 In everyday settings, behavioral contagion is manifested as the spontaneous spreading of group behavior that serves to coordinate affect and affiliation, playing a vital role in establishing group identity and co-regulation in settings such as sporting events, religious ceremonies, warfare, and rock concerts.17 Notes 1. Keysers (2011) further notes that “just as people can activate their premotor cortex without moving their own hands when seeing someone else grasp a ball, they can also activate these higher order motor areas when viewing facial expressions without necessarily moving their face” (p. 115). 2. Just as premotor mirror neurons discharge to both the sight and the sound of an agent’s actions, the insula responds both to the sight of facial expressions denoting affective states and to sounds denoting affective states. 3. For additional research on acoustic signatures see Hammerschmidt and Jürgens (2007); Juslin and Laukka (2003); Owren and Bachorowski (2007). 4. For a systematic review addressing the ecological and evolutionary bases for sound production see Eitan (2013). 5. Huron et al. (2006) draw on an ethological perspective to explore this phenomenon. In their study, listeners judged a melody in the low register with loud volume as denoting aggression or threat, and the same melody in the high register and with soft volume as denoting friendliness. Eitan and Granot (2006) also found a strong correlation between visuospatial features (size, shape, and height) and musical parameters (dynamics, pitch contour, pitch intervals, attack rate, and articulation). 6. Our human tendency to integrate visual and aural features appears to be innate (or at least prelinguistic). For additional insights on the integration of visual and aural perception see appendix V in this volume. 7. Whereas proprioception is the conscious perception of our body image, the mechanism described here circumvents any conscious processes. Proprioception, therefore, does not play a role in this model of musical empathy. For insights on proprioception and music see Acitores (2011). 8. Two studies laid the groundwork for these discoveries: Gallese et al. (1996); Rizzolatti et al. (1996). 9. For experimental research using non-invasive techniques to identify a mirror neuron system in humans see Aziz-Zadeh et al. (2004); Buccino et al. (2005); Iacoboni et al. (1999). 10. Given this shortcoming, the notion of a homologous mirror neuron system in humans has triggered much criticism. A relatively recent study using single-cell recordings in humans, however, purports to have found direct evidence of mirror neurons in humans (Mukamel et al., 2010).

Appendix I: Empathy

171

11. Studies, both in primates and humans, suggest that the sight of actions (Buccino et al., 2004), the sight of facial gestures (Dimberg et al., 2000), and the sound of vocalizations (Neumann & Strack, 2000) cause an automatic and subliminal activation of the neural structures involved in our own execution of those actions, facial gestures, and vocalizations. Lima et al. (2016) offer supporting evidence for the spontaneous activation of sound-related motor representations during auditory perception. 12. Several studies have shown that we activate regions in the insula both when experiencing an emotion or sensation and when viewing others experiencing an emotion or sensation; these findings contribute to constructing a neural representation of shared affective experiences (Carr et al., 2003). These mirroring mechanisms in the insula have been observed in relation to the perception of disgust (Wicker et al., 2003), pain (Singer et al., 2004) and touch (Molenberghs et al., 2012). 13. These mechanisms, which we develop relatively early in life, take place below the level of consciousness and are involuntary. At a musculoskeletal level, experiments using throat-audio and larynx-electromyography (EMG), show that auditory stimuli (particularly voices and music) activate subvocal articulation, even when not actively attending or aware of the stimulus (Brodsky et al., 2008). And at a more integrated (neural/embodied) level, experiments using behavioral responses (e.g., finger tapping) as well as neuroimaging (fMRI), show that our brains and bodies subliminally synchronize to periodic rhythmic stimuli (Mayville et al., 2002). Although actual movement or emotional expressions remain generally covert, in cases where inhibitory mechanisms are damaged, movement or vocalizations may manifest overtly. Numerous studies report of individuals with prefrontal cortical lesions who are unable to engage this inhibitory mechanism, resulting in unintentional and overt copying of other’s actions—this disorder is labeled as ‘echopraxia’ in psychiatry. Kinsbourne (2005) notes, however, that only actions that attract attention and provoke arousal trigger echopraxia; for instance, “[while] babies do imitate simple facial gestures . . . [they] do not imitate every movement that happens around them” (p. 166). 14. While early research framed behavioral contagion in terms of deindividuation (Festinger et al., 1952) and release restraint (Wheeler, 1966), more recent studies link behavioral contagion to the human MNS (Lee & Tsai, 2010). To date, however, there is relatively little research on the neural substrates of behavioral contagion. 15. This balance between subliminal mirroring and inhibition in empathy and social cognition is thus critical for ingroup affiliation and communication between conspecifics (Juslin & Västfjäll, 2008; Lakin et al., 2003). 16. Milgram et al. (1969) observed a correlation between the size of a crowd standing on a street looking up and the number of passersby who adopted a crowd behavior and looked up. 17. For further insights on behavioral contagion related to applauding see Freedman et al. (1980), for gaze-following see Milgram et al. (1969), and for singing along see Bargh and Chartrand (1999).

APPENDIX II Conceptual Metaphor and Image Schema

Metaphors are ubiquitous in all types of discourse, from everyday expressions to scholarly prose, from technical manuals to multimedia advertisements. They are persuasive rhetorical devices with explanatory and expressive power—explanatory because they help us grasp abstract ideas through concrete experiences, expressive because they capture and highlight objective reality while leaving ample space for creative and subjective interpretations. In this appendix, we first explore the notion of metaphor through the lens of Lakoff and Johnson’s Conceptual Metaphor Theory. Then we examine the role of image schemas within metaphorical mappings and attend more in detail to three selected schemas—CONTAINER, LINEARITY, SOURCE-PATH-GOAL—delving into their spatial logic, cultural variability, and usage within musical discourse and scholarship. Insights from experimental research in cognitive psychology and neuropsychology interspersed through the discussion validate these ideas from a scientific perspective and support the framework for meaning construction through film music I propose in Chapters 2, 3, and 4. Metaphor

In its most basic form, a metaphor is a false categorization. We recognize the statement “John is a tiger” as a metaphor because ‘John’ is not a member of the ‘tiger’ category. Understanding the meaning of this metaphor, however, is contingent upon our ability to recognize correspondences between its components—‘John’ and ‘tiger’—and map salient features between them: aggressive, untamed, or ferocious behavior.1 Figure II.1 outlines this essential cognitive mechanism of metaphor interpretation. Metaphors are conceptual and socio-cultural phenomena not bound by language, and they can, and often do, highlight correspondences between various sensory modalities.2 This led Lakoff and Johnson (1980) to suggest that metaphorical thinking is foundational to human thought and to “a more inclusive theory of human representations of reality” (p. 26). However, only when such correspondences extend to suggest semantic entailments may we fully capture a metaphor’s expressive potential.3 In Köhler (1929), we find one of the earliest references to a shape-sound correspondence and a discussion of its semantic implications.4

Appendix II: Conceptual Metaphor and Image Schema

FIGURE II.1

Metaphorical mappings.

FIGURE II.2

‘Takete’ and ‘maluma’. (After Köhler, 1929)

173

He presented participants with shapes similar to those in Figure II.2 and noted that they systematically associated angular shapes with ‘takete’ and round shapes with ‘maluma’, nonsense words projecting distinct sonic characteristics.5 Conceptual Metaphor Theory

At the core of Lakoff and Johnson’s (1980) theory is the hierarchical clustering of individual metaphors into broad categories called conceptual metaphors. For example, the metaphors “I’ve invested a lot of time”, “You are wasting my time”, and “How do you spend your time these days?” belong to the [TIME] IS [MONEY] conceptual metaphor, conveying the notion of time in terms of the function, properties, and value of money.6 Analogously, the metaphors “Mathematics has many branches”, “His ideas have finally come to fruition”, and “She has a fertile imagination” belong to the [IDEAS] ARE [PLANTS] conceptual metaphor, which frames ideas in terms of a plant’s physical structure, life cycle, and ability to reproduce organically. Clustering single examples into conceptual metaphors and expressing them in the [A] IS [B] format, as shown in Figure II.3, helps articulate the associated conceptual domains and foregrounds the embedded directionality of mappings—that is, from [B] to [A], or from ‘source’ to ‘target’.

174

Appendix II: Conceptual Metaphor and Image Schema

FIGURE II.3

Conceptual metaphor structure.

Directionality of Mappings

Kövecses (2002) argues that the source-to-target ‘unidirectional’ mapping stems from unequal degrees of abstraction between conceptual domains: “Our experiences with the physical world serve as natural and logical foundations for the comprehension of more abstract domains. This explains why in most cases of everyday metaphors the source and target are not reversible” (p. 6). Some scholars, however, challenge such unidirectionality in metaphors, arguing that in metaphors comparing two concrete or two abstract domains, the order of domains defines the directionality and hence the mapped features.7 For instance, the metaphors “That surgeon is a butcher” and “That butcher is a surgeon” compare equally concrete domains; the altered order (and implied directionality), however, drastically changes the projected features from source to target—that is, the former metaphor generates mappings of ‘careless’ and ‘brutal’, while the latter generates mappings of ‘precise’ and ‘skillful’. Multiple Sources for a Target, Multiple Targets for a Source

In constructing metaphors, we often draw on various sources for a target. For example, the target HAPPINESS enjoys various sources, such as LIGHT or FLUID IN A CONTAINER, as manifested in the metaphors “She was shining with joy” ([HAPPINESS] IS [LIGHT]) and “She overflowed with joy” ([HAPPINESS] IS [FLUID IN A CONTAINER]). Different sources depict or highlight different aspects of the target—while LIGHT attends to the energy that seems to emanate from a happy individual, FLUID IN A CONTAINER attends to the intensity and control aspects of being happy. Conversely, a single source may highlight related characteristics in multiple targets. For example, the source BUILDINGS appears in metaphors for targets as diverse as THEORIES or RELATIONSHIPS, as manifested in “His theories have a strong foundation” ([THEORIES] ARE [BUILDINGS]) and “We are building our future together” ([RELATIONSHIPS] ARE [BUILDINGS]).8 Origin and Universality of Metaphorical Mappings

Some scholars trace the origin of metaphorical mappings to primal, body-based experiences (e.g., Grady, 1997; Lakoff, 2014).9 For instance, the conceptual metaphor [AFFECTION] IS [TEMPERATURE], manifested in “He is a warm person” or “She gave him the cold shoulder”, arguably emerge from a “systematic co-occurrence of feelings of physical warmth and feelings of affection in earliest infancy” (Bundgaard, 2009, p. 402).10 Bargh et al. (2010), in turn,

Appendix II: Conceptual Metaphor and Image Schema

175

primed participants with physical experiences of warmth, distance, hardness, weight, and roughness, and found that such priming strategies strongly influence social judgments, leading them to advance the notion of “a cognitive architecture in which social-psychological concepts metaphorically related to physical-sensory concepts—such as a warm person, a close relationship, and a hard negotiator—are grounded in those physical concepts, such that activation of the physical version also activates (primes) the more abstract psychological concept” (p. 268).11 Kövecses (2002) offers a broader perspective, tracing the origin of mappings to perceptual, biological, or cultural experiences: [MORE] IS [UP] (“Turn the radio down”) stems from correlations of perceptual experiences, such as adding fluid to a container and observing its level rise; [ANGER] IS [HEAT] (“I am boiling with anger”), stems from correlations of biological experience, such as feeling internal temperature changes resulting from mood changes; and [LIFE] IS [A GAMBLING GAME] (“It is a roll of the dice”) stems from correlations of cultural experiences, such as defining our actions as social activities that involve luck and randomness. This broader perspective regarding the origin of metaphors foregrounds that cultural and linguistic variations often highlight a tension between physically and culturally grounded metaphors. To resolve this tension, Ibarretxe-Antuñano (2013) proposes the ‘culture sieve’ model, a two-stage process in which “bodily-grounded experiences that contribute to understanding and motivating the metaphorical mappings between the source and target” are subsequently “purged, adapted, and modified by the cultural information available” (p. 324).12 Aware of this tension between universality and cultural variation, Dewell (2005) warns us that although “semantic meanings are grounded in prelinguistic cognition that is real and patterned and largely universal” we must avoid “the implication that the meaning of language constructions can be reduced to pre-existing universal representations” (p. 385). Image Schemas

Image schemas are integral to metaphorical constructions. These emerge from recurrent embodied interactions with the environment, capturing regularities in our sensorimotor experiences as spatial and bodily logics—placing objects within a CONTAINER, traversing a LANDSCAPE from a SOURCE through a PATH and toward a GOAL. Schemas thus become mental constructs that enable metaphorical mappings, thereby shaping our perception, discourse, and even our abstract thoughts.13 For example, the conceptual metaphor [RELATIONSHIPS] ARE [JOURNEYS]—instantiated in “Marriage is a long and bumpy road”, “Our relationship is moving along in the right direction”, and “We may have to go our separate ways”—activates the SOURCE-PATH-GOAL image schema and serves as a vehicle to understand the abstract notion of relationships in terms of our concrete experiences of journeys. Given the abstract nature of music, it is not surprising that we conceptualize it using metaphors that draw on image schemas.14 In fact, descriptions of musical phenomena activate many image schemas, such as in “The clarinet enters [CONTAINER, PATH] the orchestral texture with a rising [VERTICALITY, LINEARITY] melody traversing [PATH] two octaves, signaling the imminent arrival [PATH-GOAL] of a closing cadence in a distant [LANDSCAPE] key area [LANDSCAPE]”. Such prose, filled with metaphors that activate image schemas, is arguably necessary to communicate and share interpretations of music.15

176

Appendix II: Conceptual Metaphor and Image Schema

CONTAINER Schema

The CONTAINER schema emerges from concrete embodied experiences with bounded regions of space, such as interacting with their contents (placing cookies in a jar, filling a cup) or crossing their boundaries (walking out of a building, diving into the ocean, getting in the car). Figure II.4 presents a diagram of the CONTAINER schema, which includes a region of space and a boundary delineating an interior and exterior. Lakoff and Johnson (1980) regard our body as the primary ‘container’, as we are “bounded and set off from the rest of the world by the surface of our skins . . . [projecting] our own in-out orientation onto other physical objects that are bounded by surfaces” (p. 29). Once we establish the CONTAINER schema via concrete embodied experiences, we overlay its spatial logic onto abstract phenomena that objectively lack spatial properties. This allows us to grasp a wide array of abstract thoughts and phenomena, from actions framed by cultural constructs (such as crossing national borders or entering a new century) to metaphorical expressions that extend beyond the empirically plausible (“My heart is full of love”, “I passed out”, “We are opening an investigation”, “I’ll keep that in mind”).16

FIGURE II.4

Diagram of the CONTAINER schema.

Although seemingly universal, the CONTAINER schema is not free from cultural variability or unfettered by the influence of language. For instance, while English speakers use ‘in’ and ‘on’ to denote spatial relationships (which draw on the CONTAINER and SUPPORT schemas correspondingly), Korean speakers use the verb kkita to denote gradations of ‘tight’ versus ‘loose’ fit to denote the same spatial relationships (Aurnague et al., 2007; Choi et al., 1999). Vandeloise (2007) draws on the CONTAINER schema to illustrate a similar phenomenon, noting that in English, blended mixtures are often (albeit erroneously) conceptualized in terms of CONTAINMENT; when using sentences such as “There is milk in the coffee”, we reinforce a false cultural frame in which the coffee, as the more important of the two liquids, is the container, and the milk, as the lesser liquid, is the contained. These examples illustrate that, because schemas ground linguistic and semantic meaning, they shape our perception of reality. LINEARITY and VERTICALITY Schemas

Concrete embodied experiences framed within a one-dimensional continuum—e.g., wherein size (small to large), height (low to high), or speed (slow to fast) play a role—bring about the LINEARITY schema. In turn, the VERTICALITY schema emerges from a subset of those experiences, where the vertical orientation (in addition to the one-dimensional continuum) is of particular relevance; examples include kinesthetic experiences of traversing vertical distances by climbing up or down (ladders, mountains), observational experiences of gazing up or down at vertical structures (columns, trees), and embodied experiences entailing the co-occurrence of events (such as pouring liquid in a container and observing its

Appendix II: Conceptual Metaphor and Image Schema

FIGURE II.5

177

Diagrams of the VERTICALITY and LINEARITY schemas.

level rise).17 As diagrammed in Figure II.5, both exhibit a one-dimensional continuum; but while VERTICALITY denotes a specific orientation in space (perpendicular to the horizon), LINEARITY lacks a specific orientation. Once established through concrete embodied experiences, the LINEARITY and VERTICALITY schemas permeate our understanding of abstract concepts by scaffolding their continuous one-dimensional structures (with or without a predetermined vertical orientation). This allows us to construct metaphors where the orientation is ambiguous or altogether absent (“He has mood swings”, which draws on LINEARITY) or where the vertical orientation plays a vital role (“The rising violence in the city”, which draws on VERTICALITY). The LINEARITY schema provides a more suitable alternative (than the VERTICALITY schema) for conceptualizing a wide range of abstract phenomena, as it allows us to reveal their onedimensional linear quality while circumventing possible orientation constraints. By drawing on concrete experiences framed by one-dimensional continua, the LINEARITY schema thus shapes our perception and cognition of abstract notions, including socio-cultural ones such as social class (low to high) and interpersonal affection (warm to cold), as well as musical ones such as pitch (low to high), tempo (slow to fast), and volume (soft to loud). SOURCE-PATH-GOAL Schema

The SOURCE-PATH-GOAL schema (SPG) emerges from concrete embodied experiences featuring a start-point, a path, and an end-point, such as going from home to work via our usual route. Its spatial logic, diagrammed in Figure II.6, comprises a starting-point (A), an endingpoint (B), a path connecting the two, and the implicit movement from A to B.18 Once established, the SPG schema provides robust explanatory power for abstract constructs featuring a start-point, a path, and an end-point. To better grasp the abstract notion of ‘relationships’, for instance, we use expressions such as “Our relationship has hit a deadend”, “We may have to go our separate ways”, “It’s been a long and winding road”, all based on the [RELATIONSHIPS] ARE [JOURNEYS] conceptual metaphor grounded in the SPG schema. Analogously, we resort to the spatial logic of the SPG schema to account for the syntagmatic unfolding of narrative events. Metaphors that describe narrative developments, for instance, include expressions such as “The characters embarked on a journey”, “They encountered obstacles along the way”, and “The protagonist became side-tracked”.

178

Appendix II: Conceptual Metaphor and Image Schema

The SPG schema often interacts with other schemas that complement its structure, resulting in multiparametric blends that more truthfully capture the essence of a metaphor’s concrete and abstract domains. For instance, although not a metaphor in itself, the English word ‘into’ simultaneously activates the CONTAINER schema (the ‘in’ portion) and the SOURCEPATH-GOAL schema (the ‘to’ portion), as diagrammed in Figure II.7. Combining the spatial logic of these two schemas allows us to describe concrete actions (walking into a building, diving into the ocean) and grasp abstract ideas (fading into oblivion, signing into your email account).

FIGURE II.6

Diagram of the SOURCE-PATH-GOAL schema.

FIGURE II.7

Diagram of the SOURCE-PATH-GOAL and CONTAINER schemas combined.

Notes 1. Reference to context is critical in identifying and understanding metaphors, as contextual clues can transform a factual statement into a metaphor and vice versa. For example, although “John is a tiger” may justifiably seem a metaphor, in the context of sports, this apparently false categorization could be a factual one—that is, that John either belongs to or roots for the Detroit Tigers (or one of the many tiger-themed sports teams). 2. Linguistic cross-modal metaphors can evoke various sensory modalities. For example, in “The dark sound of thunder”, the semantic sphere of ‘dark’ activates our sight while ‘sound of thunder’ activates our hearing.

Appendix II: Conceptual Metaphor and Image Schema

179

3. Our ecological and evolutionary biology drives our tendency to attend to congruent stimuli across senses. Bregman (1990) identifies this tendency in newborns, who “spend more time looking at a face that appears visually to be speaking the same words that it is hearing than one that is not”, suggesting this is evidence that “the grouping of sounds can influence the grouping of visual events with which they are synchronized and vice versa” (p. 653). Similarly, Welch’s (1999) ‘unity assumption’ describes the human tendency to attend to structurally congruent stimuli; he speculates that humans are “vested in maintaining congruence in their perceptual world” (p. 373) and hence in all modes of perception. Related to film music, Marshall and Cohen’s (1988) study, particularly their Congruence-Association Model, supports the claim that “attention is directed to the overlapping congruent meaning of the music and film” (p. 109); although this study has been widely cited in the film music literature, much research exploring cross-modal and cross-domain correspondences using innovative (and more robust) experimental paradigms have challenged its findings (e.g., Lipscomb, 1995; Millet et al., 2021; Sirius & Clarke, 1994). Moreover, many studies identify or corroborate cross-domain correspondences without delving into the potential semantic entailments of these correspondences. Velasco et al. (2016) tackle taste-vision correspondences and examine the extent to which a food package’s shape (straight, curvy, symmetrical, asymmetrical) influences the consumers’ taste perceptions and expectations; their results broadly suggest that packages with round contours induce expectations of sweet flavors, while sharp edges induce expectations of bitter flavors. Sweeny et al. (2012) focus on visual-aural correspondences that stem from systematically seeing mouth shapes (vertically versus horizontally stretched) and hearing particular sounds (‘wee’ versus ‘woo’); their results complement the McGurk effect (where what we hear is shaped by what we see) by suggesting the opposite phenomenon, that “speech sounds influence visual perception of a basic geometric feature . . . what we see is shaped by what we hear” (p. 199). 4. When identifying and corroborating cross-modal and intra-modal correspondences, empirical and experimental studies in cognitive musicology echo the studies (and results) from the broader cognitive sciences (e.g., Collier & Hubbard, 2001; Eitan & Granot, 2006; Gjerdingen, 1994; Lipscomb & Kendall, 1994). Music cognition studies that venture into the neurosciences also echo the corresponding claims and results, suggesting that conceptualizing music and musical meaning recruits somatomotor and somatosensory cortical areas (e.g., Platel et al., 1997; Zatorre et al., 1999). Most studies in music cognition, however, explore these phenomena by presenting participants with stimuli that isolate a single musical or sonic parameter (such as single notes, scales, beats), stimuli best characterized as ‘sounds’ rather than ‘music’; nevertheless, this body of research has been the starting point of a journey leading to the theories I put forth in this volume. 5. Many studies extend Köhler’s work, tracing the origin of the cross-modal correspondences he identified while speculating about their universality, ontogeny, and evolutionary implications. Several studies argue that this cross-modal correspondence rests on the covert activation of vocal articulatory movements (e.g., Deroy & Auvray, 2013; Galantucci et al., 2006). On the other hand, Ozturk et al. (2013) replicated Köhler’s study with infants, using the words ‘kiki’ and ‘bubu’, and their results suggest that such associations are preverbal. In exploring the cultural variability of this cross-modal correspondence, Ramachandran and Hubbard (2001) also replicated Köhler’s experiment, this time with American and Indian participants and using the words ‘kiki’ and ‘bouba’; their results suggest that such shape-sound correspondences are to some degree cross-cultural. Bremner et al. (2013) further elaborate on the cross-cultural pervasiveness of this correlation. 6. The use of capitals for conceptual metaphors and image schemas is standard practice within this field of scholarship; the use of brackets for target and source domains is not, but it helps delineate the components of a conceptual metaphor. 7. Ortony (1993), for instance, argues that the mapped features define the directionality, because these are salient in the source, not the target. 8. Bundgaard (2009) notes that some abstract targets are consistently, and at times exclusively, framed through single concrete sources—for instance, the target TIME is generally framed by the source SPACE. 9. Johnson (2007) suggests that both domains entering a metaphor, the concrete and the abstract, share neural substrates, that “the body must recruit neural structures central to sensory and motor processing to carry out the inferences that make up our abstract patterns of thinking. Structures of perceiving and doing must serve as structures of thinking and knowing” (p. 26, emphasis in original). To support this argument, some scholars recur to computational models of neural networks (Feldman & Narayanan, 2004); such models, however, redirect the conversation away from

180

10.

11.

12.

13.

14.

15. 16. 17.

18.

Appendix II: Conceptual Metaphor and Image Schema

embodiment and toward logical and computational paradigms typical of the first wave of cognitive psychology. Such cross-domain co-activations at the early stages of life arguably form the basis of conceptual metaphors such as [DESIRE] IS [HUNGER] (“She is starving for attention”). Johnson and Lakoff (2002) note that “we first acquire the bodily and spatial understanding of concepts and later understand their metaphorical extensions in abstract concepts” (p. 254). Asch and Nerlove (1960) provide evidence consistent with Johnson and Lakoff’s claim; in their study, children initially developed a command of words like ‘warm’ or ‘cold’ within the context of objects, and only later applied these concepts to individuals using qualifiers such as ‘warm’ or ‘cold’ to describe their personality. Although insufficient to draw conclusive results, a growing body of research suggests that humans recruit sensorimotor brain areas for abstract reasoning and semantic comprehension. For instance, using event-related fMRI, Hauk et al. (2004) found that action words associated with particular body parts (‘lick’, ‘pick’, ‘kick’) activate areas in the premotor cortex directly adjacent to (or overlapping) the areas activated by the movement of the corresponding body parts. More recent work within cognitive neuropsychology merging theoretical, meta-analytical, and computational accounts explains this phenomenon as a co-activation or re-activation of neural substrates during metaphor or image schema processing. For instance, Anderson’s (2010) ‘neural reuse’ theory, which draws on a database of thousands of fMRI studies, proposes that both domains in a metaphor (the abstract and the concrete) share neural substrates. He argues that “conceptual-linguistic understanding might involve the reactivation of perceptuomotor experiences . . . [in metaphorical mappings] understanding in one domain would involve the reactivation of neural structures used for another” (p. 254)—in other words, when unpacking a metaphor to grasp abstract ideas, we must reactivate neural structures previously developed for concrete experiences. Abril (2001) studied how bilingual children discern musical registers, noting English terminology uses ‘high/low’—i.e., drawing on the VERTICALITY schema—while Spanish terminology uses ‘agudo’ (sharp) and ‘grave’ (serious)—i.e., deviating from both VERTICALITY and LINEARITY schemas. Talmy (1983) and Langacker (1987) were among the first to put forth the notion of image schema. Johnson (1987) reformulated the notion of image schema as “a dynamic cognitive construct that functions somewhat like an abstract structure of an image and thereby connects together a vast range of different experiences that manifest this same recurring experience” (p. 2). Here, I use the terms ‘schema’, ‘embodied schema’, and ‘image schema’ interchangeably, and use ‘schemas’ instead of ‘schemata’ to denote the plural. Many musicologists and music theorists have drawn on the notions of conceptual metaphor and image schema to formulate broad frameworks for musical thinking and musical meaning (e.g., Chattah, 2006, 2015; Cook, 1998; Cox, 1999, 2016; Johnson & Larson, 2003; Larson, 2012; Saslaw, 1996; Spitzer, 2004; Zbikowski, 1998, 2002). Larson (2012), for instance, focuses on musical ‘forces’ such as ‘gravity’, ‘magnetism’, and ‘inertia’, understood as recurrent bodily experiences of motion. Cox’s (2016) ‘mimetic hypothesis’ takes conceptual metaphor as a point of departure in recognizing the “role of physical imitation in everyday cognition, and on the various forms of physical imitation in music cognition” (p. 3). And Zbikowski (2002), perhaps the most significant contribution toward systematizing the music-theoretical understanding of metaphor, extends cross-domain mappings to include Fauconnier and Turner’s (2002) Conceptual Blending theory, which we further explore in Appendix VII in this volume. Just like metaphors, schemas are not bound by language. For instance, the various diagrams included in this and other chapters (single lines, arrows, cubes) activate the very schemas they represent. While schemas emerge from recurrent embodied experiences, cultural practices often frame these experiences. At the neural level, establishing the VERTICALITY schema may entail the simultaneous neural activation of areas dedicated to two unrelated phenomena (such as pouring liquid in a container and observing its level rise) where “experiential conflation has no semantic motivation and is solely identified as simultaneous activation of distinct parts of the brain. Frames or domains experienced together are temporally neurally bound: they fire in synch” (Brandt, 2009, p. 66). The SPG schema is analogous to the LINEARITY schema in its one-dimensional structure, while additionally featuring a defined orientation, from ‘source’ to ‘goal’.

APPENDIX III Affordances

James Gibson challenged the prevailing paradigms in psychology, which held that our minds construct representations of the world around us and that perception is a passive and receptive mechanism unaffected by our environments. He advanced an ecological approach to psychology, arguing that perception does not rely on cognitive or symbolic constructions and that our senses develop functional adaptations in response to changes in our environments. Gibson’s viewpoints and arguments, particularly his notion of ‘affordances’, rapidly attracted the attention of researchers in the cognitive and social sciences, who embarked on an empirical exploration of his novel conceptual frameworks. The findings from these studies have in turn informed recent advances within widely diverse areas of knowledge, including artificial intelligence, robotics, industrial design, human factors, architecture, and the arts. In this appendix, I briefly expound on the notion of affordances, summarize relevant insights from the cognitive sciences, and survey its recent application within music-related scholarship. In Chapter 5, I further extend Gibson’s argument by claiming that film music affordances not only allow but also invite a wide array of potential (inter)actions—during a film, we ‘resonate’ with the sonic environment, attuning our perceptual experiences to the music’s affordances, and in turn, these latent or actualized affordances influence our bodily responses and guide our interpretations of the narrative. Theory of Affordances

An ecological account of perception rejects the notion that our senses passively transform information from the outside world into neural patterns; instead, it contends that our senses actively detect and interpret ecologically relevant information—we do not simply see; we recognize opportunities for (inter)action. Gibson (1979) coined the notion of ‘affordances’ to denote the potential (inter)actions the environment offers us—a chair affords us something to sit on, an enclosed space affords us shelter, a bridge affords us an opportunity to traverse a distance between two points.1 Affordances combine perception and action into a unified and interdependent process. Perceiving potential (inter)actions means perceiving affordances—the kick-ability of a ball,

182

Appendix III: Affordances

the grasp-ability of a mug, the climb-ability of a hill.2 Because such potential (inter)actions rest on a reciprocal relationship between an organism’s capacities and its environment’s material properties, affordances are relative to and co-specified by both these variables. For instance, a body of water affords a support surface to a water strider bug, but not a human. Furthermore, despite our environments’ relatively stable structural properties and our relatively stable bodily capacities, new (and creative) affordances emerge based on our needs, goals, or intentions.3 As a result, affordances are in constant flux—although a chair affords us to sit on it, the same chair may afford us to hang our coats, fight an intruder, lock a door, or step on it to reach a ceiling fixture. In advancing and refining his ecological approach to cognition, Gibson (1979) argues that because we attune our perception so strongly to our intended motor actions, “what we perceive when we look at objects are their affordances, not their qualities” (p. 134). This means that the environment’s affordances determine the attributes we perceive in it, that functional significance (rather than material structure) shapes our perception. For instance, we perceive a chair not as an object of a particular color or shape but as an object that affords us something to sit on (or do something else with).4 Such interdependency of perception and action has inspired a wealth of empirical studies within cognitive psychology and the neurosciences.5 Empirical Explorations of Affordances

The notion of affordances has received ample empirical support from various research avenues, each focusing on different facets. While studies in cognitive psychology have sought to understand human behavior in relation to the environment’s affordances, studies in neuroscience have sought to find neural representations of affordances. Perspectives from Cognitive Psychology

Most studies in cognitive psychology explore human behavior as influenced by the affordances of objects and surfaces.6 For instance, Tucker and Ellis (1998) investigate the effect of an object’s affordance on our response times. In one of their experiments, they presented photographs of graspable objects to participants and recorded their reaction times and response mode. When the graspable portion of the object was oriented to the right, right-hand responses were faster; when oriented to the left, left-hand responses were faster. These results support the proposition that objects within our peripersonal space “automatically potentiate components of the actions they afford” (p. 830). Studies exploring human behavior in relation to the affordances of surfaces support Gibson’s (1979) idea that “the composition and layout of surfaces constitute what they afford . . . to perceive them is to perceive what they afford” (p. 127). For instance, Adolph et al. (1993) explore the perception of the traversability of sloped surfaces; in their study, toddlers altered their locomotor strategies to the specific affordances of various sloped surfaces (by walking, crawling, sliding, and even refusing to traverse). Similarly, Warren (1984) presented different riser heights to participants and recorded their behavior; their results support the proposition that our environment’s affordances guide our behavior, that “by perceiving environmental objects in relation to their action capabilities, [we] can detect the possible and most economical courses of action” (p. 700).

Appendix III: Affordances

183

Perspectives from Neuroscience

Gibson’s intuitions anticipated recent neuroscientific findings. At a time when neurons were believed to fire during either perception or action, Gibson suggested the presence of bimodal neurons that fire during both perception and action. As early as 1966, he noted that “[t]he individual nerve or neuron changes function completely when incorporated in a different system or subsystem” (p. 56); and in 1979, he stated that organisms have “subsystems for perceiving the environment and concurrently for getting about in it and manipulating it” (p. 225). Through the notion of ‘subsystems’, Gibson dissolves the distinction between perception and action and anticipates one of the most groundbreaking advances in neuroscience: the discovery of canonical neurons.7 Canonical neurons are bimodal neurons that fire both when observing the object and when actively interacting with it. These neurons “respond selectively to the presentation of a three-dimensional object, in function of its shape, size, and spatial orientation . . . [and] in terms of the type of interaction they allow” (Garbarini & Adenzato, 2004, p. 102); this means that they respond not only to the material properties of an object, but to its potential affordances.8 Studies that support the presence of canonical neurons have focused on the co-activation of ‘motor control system’ areas in the brain; for example, when presented with a graspable object, canonical neurons fire both when observing and when grasping the object. These findings further suggest that affordances result from direct perception (not from cognitive construction), emerge from below the level of consciousness, and do not rely on symbolic representations.9 Many studies have contributed to formulating a Gibsonian neuroscience with various degrees of success, most of them proposing that “the primary function of perception is not to represent the world, but to guide an animal’s actions” (de Wit et al., 2017). Some studies have extended the notion of affordances, reexamining the interplay between the environment’s structure, our bodily capacities, and our intentions, all from the perspective of neural processing. For instance, Ellis and Tucker (2000) suggest that “seen objects potentiate a range of actions associated with them, irrespective of the intentions of the viewer” (p. 451); they refer to these as “micro-affordances”, which are present in the nervous system as “dispositional states”. Using functional magnetic resonance imaging (fMRI), Anderson (2016) identified flexible structural-functional connections in the brain; he noted that: many of the properties to which the brain is attuned are likely to be action-relevant and relational . . . Interaction with an environment offering multiple affordances causes regions of the brain to be differentially activated in accordance with their functional biases. (p. 8) Moreover, Cisek’s (2007) ‘affordance competition hypothesis’ suggests that our brains simultaneously process sensory, motor, and cognitive information provided by regions in each of the four lobes of the cortex and that “sensory information arriving from the world is continuously used to specify several currently available potential actions” (p. 1586). Cisek’s hypothesis thus supports the notion of affordances as ‘latent’ action potentials that “may or may not be actualized on a given occasion, but it nonetheless represents a ‘real possibility’ of action” (Pinna, 2017).

184

Appendix III: Affordances

Extended Affordances

Gibson (1979) reminds us that our environments afford us “a rich and complex set of interactions, sexual, predatory, nurturing, fighting, playing, cooperating, and communicating” (p. 128), indicating that affordances extend beyond the physical or material and onto the social and intellectual realms. Even the material properties of our environments offer these more abstract affordances. For instance, while we mainly use our hands and fingers to manipulate objects, we also often use our fingers to perform basic mathematical operations.10 Our understanding of ‘environment’, however, should extend beyond material or physical attributes. Gibson (1979) remarks that: it is a mistake to separate the natural from the artificial as if there were two environments . . . to separate the cultural environment from the natural environment, as if there were a world of mental products distinct from the world of material products. (p. 130) Such a broad understanding of the notion of ‘environment’ allows us to identify potential affordances that extend beyond the bodily or physical realms and into abstract and intellectual ones. We live in environments saturated with cultural and social constructs; language, for instance, affords us communication, discussion, persuasion, confrontation, cooperation, expression, comprehension, representation, description, deception, translation, and so much more. Art is an integral part of our cultural and social environments, one that initially engages us physically but also provides opportunities for widely diverse kinds of intellectual engagement. Music, especially, is a medium that illustrates a broad range of affordances, from physical and physiological (dance, play, exercise, mood modulation) to social and psychological (communication, political expression, identity construction).

Music’s Affordances

Music’s affordances emerge from an interplay among three conditions: the music’s structure, our perceptual capacities, and our intentions. In terms of the music’s structure, the vast number of musical parameters (e.g., timbre, dynamics, meter, harmony, articulation, density, texture) bring about an almost limitless set of affordances. In terms of our perceptual capacities and intentions, Clarke (2005) reminds us of the wide variety of potential affordances of music, which include “dancing, singing (and singing along), playing (and playing along), working, persuading, drinking and eating, doing aerobics, taking drugs, playing air guitar, traveling, protesting, seducing, waiting on the telephone, sleeping . . . the list is endless” (p. 204). The last decade has seen an outpouring of scholarly explorations of music’s affordances; these range widely in scope and perspective, accounting for our physical engagement with music (performing, dancing, exercising) as well as our intellectual engagement with it (analyzing, communicating, signifying).11

Appendix III: Affordances

185

From a performative perspective, Huron and Berec (2009) explore how music affords instrumentally idiomatic gestures; they note that: the properties of musical instruments . . . suggest certain modes of interaction or encourage certain musical functions . . . A stretched-membrane drum, for example, affords many sonic possibilities . . . [it] may be struck with a hand or with a stick; it may be struck by a single hand, by alternating hands, by drumming the fingers. (p. 104)12 From a similar performance practice perspective, Love (2017) explores how stylistic norms define the musical environment and constrain the potential affordances; for example, when improvising within the jazz style, the performer “perceives this environment in terms of its affordances . . . [and] seeks opportunities for artistic display” (p. 31). Krueger’s (2011) discussion of music’s affordances includes the potential for emotion regulation, communicative expression, and identity construction. He suggests that music affords us opportunities for: creating and cultivating the self, as well as creating and cultivating a shared world that we inhabit with others . . . Thinking of music as an affordance-laden structure thus reaffirms the crucial role that music plays in constructing and regulating emotional and social experiences in everyday life. (p. 1) Clarke (2005), in turn, redirects the conversation toward an intellectual engagement with music and fully frames the notion of affordances as the potential for analytical interpretation and musical criticism. In exemplifying such an account of music’s affordances, he notes that: in the specific contexts of musical hermeneutics, musical material can be conceived as affording certain kinds of interpretation and not others . . . The recapitulation of the first movement of Beethoven’s Ninth Symphony affords writing (or speaking) about in terms of murderous sexual rage, or the heavens on fire. (p. 203) Notes 1. The ‘material’ environment is here understood as consisting of all surrounding objects, spaces, and surfaces. 2. Gibson’s ecological approach aligns with the notion of an ‘embedded mind’ not dissociated from natural, social, and cultural environments. 3. Norman’s (2002) “physical versus learned” categorization of affordances acquires particular relevance in light of the current trends of “life hacks”, which entails the novel use of familiar artifacts to arrive at practical solutions to common problems. The emergence of life hacks further underscores that the object-perceiver relationship is centered on, and relative to, an organism’s intended actions. 4. Implicit in the notion of affordances is a critique of the Cartesian mind-body dualism. Gibson (1979) suggests that perception does not elicit or give rise to neural representations of the account of perception—that is, perception entails ‘resonating’ with the environment’s affordances, and further notes that “exteroception is accompanied by proprioception . . . to perceive the world is

186

5.

6. 7. 8. 9. 10.

11.

12.

Appendix III: Affordances

to coperceive oneself . . . The awareness of the world and of one’s complementary relations to the world are not separable” (Gibson, 1979, p. 141). Research in cognitive psychology and neuroscience, however, faces methodological and ontological challenges in collecting and interpreting data to support Gibson’s notion of affordances. In fact, a neural account of perception and action seemingly counters Gibson’s ecological account of perception–that is, locating affordances as representations inside the brain conflicts with an ecological program by addressing the very subject-object framework Gibson attempted to disprove. For example, many studies in cognitive psychology deviate from Gibson’s ecological approach by reinstating the subject-object and mind-body dualist views, framing affordances as the neural ‘representations’ or dispositions of motor actions (Declerck, 2013; Sakreida et al., 2016). For studies in cognitive psychology that explore the affordances of tools see Bach et al. (2014); Fagg and Arbib (1998); Kühn et al. (2014); Makris et al. (2013); Proverbio et al. (2013); Sakreida et al. (2016). For further insights on canonical neurons (which are closely related to mirror neurons) see Iacoboni et al. (2005); Rizzolatti and Fadiga (1998); Rochat et al. (2010). Another type of bimodal neurons, mirror neurons, fire both when performing an action and when seeing or hearing someone else performing that same action. These bimodal neurons are foundational to the model of musical empathy proposed in Chapter 1. Such co-activations extend to aural perception (Osiurak et al., 2017). For instance, Chao and Martin (2000) found evidence that seeing or reading the names of tools or manipulable artifacts activates brain areas involved in grasping. This type of cognitive off-loading is just one instance in which we use the environment’s affordances to engage in more abstract behavior. Such cognitive off-loading by using “the structure of the environment and their operations upon it as a convenient stand-in for the information-processing operations” (Clark, 1989, p. 64) has been observed exclusively in humans. Reybrouck (2005) devises a taxonomy of music’s affordances. He introduces five categories of affordances: “[i] the sound-producing actions proper, [ii] the effects of these actions, [iii] the possibility of imagining the sonorous unfolding as a kind of movement through time, [iv] the mental simulation of this movement in terms of bodily based image schemata and [v] the movements which can be possibly induced by the sounds” (pp. 258–259). In a subsequent study (2012), he adds three related categories of affordances related to the performative aspects of music: “(i) the production of musical instruments out of sounding material, (ii) the development of playing techniques to produce musical sounds, and (iii) the shaping of the sound by using modulatory techniques” (p. 404). For additional insights on instrumental affordances see De Souza (2017).

APPENDIX IV Memory

Memory intersects with nearly every component of our human experience, from engaging in physical activities (riding a bike) to communicating (language) and even developing a personal or collective identity. And although it seems paradoxical, while our memories are grounded in the past, one of their primary roles is to predict and anticipate the future. This appendix first surveys well-established theories about memory systems and then explores current theories explaining memory’s ecological and evolutionary functions. Memory Systems Theories

Recognition, recollection, retrieval, remembering, knowing, familiarity, all involve various levels of engagement with information and are cast under the broader umbrella of memory. While some scholars frame the various levels of engagement with information within hierarchically arranged category structures, others characterize these as graded manifestations of a single phenomenon within a continuum. This has opened the space for multiple, often competing yet overlapping theories of memory systems which describe similar concepts using identical terminology. In this first part, we survey three theories: record-keeping, working memory, and long-term memory. Record-Keeping

The record-keeping theory is what most of us think of as memory. Via the ‘storage’ metaphor, this theory likens human memory to that of machines or recording devices.1 It proposes a memory architecture that features three types of storage—sensory, short-term, and longterm—each defined in terms of its modality, capacity, and duration (Atkinson & Shiffrin, 1968). Sensory memory is modality-specific, holding a finite amount of perceptual information that we may retrieve for a few seconds after exposure. In the auditory domain, this is called ‘echoic memory’, where aural stimuli may be ‘played back’ in our minds with relatively good fidelity for a short time (Winkler & Cowan, 2005).2 After a few seconds, our sensory

188

Appendix IV: Memory

memories fade; this information decay stems from (1) an extended time-lapse to a repetition of the stimuli or (2) interference of competing information (Berman et al., 2009).3 Two other types of information storage within the record-keeping theory are ‘short-term’ and ‘long-term’ memory, each defined in terms of their temporal reach and capacity. Such a division between short-term and long-term memory as separate storage locations aligns with the memory architecture of computers as having separate short-term and long-term memory modules. While short-term memory holds a limited amount of information for a limited time, long-term memory holds vast amounts of information that we can retrieve or access as needed, even years later.4 Given the difficulty in adequately explaining the process whereby we consolidate sensory memories into short-term ones, and short-term memories into long-term ones, current theories reject the idea of such processes altogether. Instead, alternative memory-systems models contend that when performing memory-related tasks, we do not draw on a single memory system but combine and integrate several of them.5 Working Memory

Baddeley and Hitch (1974) challenged the (then-dominant) model of short-term memory as temporary information storage, proposing a set of real-time reconstruction processes instead. In their “working memory” model, information from sensory memory is actively (and often subliminally) manipulated via complex cognitive tasks (e.g., retrieving longterm memory frames) to consolidate long-term memories.6 Therefore, working memory rests on a subliminal interaction between sensory and long-term memories through which we assimilate new stimuli into existing knowledge and thereby consolidate long-term memories.7 Several mechanisms are vital to working memory: repeating (e.g., phonological loop), chunking (based on gestalt principles), recognizing (based on schemata), activating (via priming and classical conditioning), and associating (via semantic and episodic memory systems).8 Repeating and chunking involve bottom-up processing, and hence are not driven by interpretative mechanisms. Repeating is critical in consolidating memories (Jacoby, 1978). For instance, the ‘phonological loop’ is an essential mechanism of auditory working memory; it engages subvocal rehearsal to keep aural information from decaying. Chunking entails integrating discrete pieces of information into single, larger units (Mathy & Feldman, 2012). For instance, when exposed to raw information, we engage gestalt principles to group that information, to ‘chunk’ it.

FIGURE IV.1

Working memory systems.

Appendix IV: Memory

189

In contrast to repeating and chunking, which involve bottom-up processing, recognizing, activating, and associating involve top-down processing that draws on long-term memory systems. Recognizing engages schemata as pre-existing templates to guide our perception and behavior.9 Activating engages classical conditioning and priming by eliciting related long-term memories; we may consider this a precursor to establishing semantic or affective associations. Associating (semantic or affective) meaning to stimuli helps consolidate memory traces; for instance, we are more likely to retain a known word such as ‘bicycle’ than a nonword that uses the same syllables such as ‘clebicy’ (Hulme et al., 1991).10 Long-Term Memory

Schacter and Tulving (1994) complicate the traditional conception of long-term memory as passive retention, identifying multiple long-term ‘memory systems’. Their primary subdivision of long-term memory systems includes implicit (or non-declarative) and explicit (or declarative) memories, each subdivided further.11 We engage our implicit memory when performing tasks that rely on habit or skill—riding a bicycle, tying shoelaces, driving home from work. Implicit memory is “encoded during a particular episode [and] subsequently expressed without conscious or deliberate recollection” (Schacter, 1987, p. 501); subsequently, it “does not involve conscious recollection but instead reveals itself through behavior” (Eysenck & Keane, 2015, p. 286). Implicit memory is understood through various systems: priming, classical conditioning, schemata, and scripts. This subdivision notwithstanding, brain research shows that all long-term memory systems are interdependent and rely on one another for consolidation and subsequent retrieval (Burianova et al., 2010). Priming and classical conditioning outline cognitive processes influenced by a previously presented stimulus. In priming, exposure to a stimulus influences our behavior, response, or perception of a subsequent stimulus. For example, primed with the color yellow and subsequently asked about fruits, most of us think of a banana or a lemon, ‘activating’ that concept because of the perceptual relationship between stimulus and response. Besides such perceptual relationships, priming paradigms may include semantic or conceptual relationships. While priming subtly influences our response or behavior, classical conditioning more directly elicits an associated response or behavior. Classical conditioning stems from establishing an association between two contiguously presented stimuli so that one evokes the

FIGURE IV.2

Long-term memory systems.

190

Appendix IV: Memory

presence of the other.12 In Pavlov’s familiar experiment, the dogs associated the sound of a bell with food; hence, the sound caused them to salivate. Soon after Pavlov published the results of his experiments, much empirical research explored classical conditioning in humans and within a wide range of settings, from restaurant aromas to routine immunizations to everyday cellphone use.13 Schemata and scripts, two other systems of implicit long-term memory, are abstract structures characterized by patterns of thought or behavior. Schemata emerge through recurrent experiences of interaction with our environments, shaping our perception and behavior (i.e., recognizing people, objects, or sounds). An experiment by Barlett (1933) illustrates the scaffolding power of schemata; in the experiment, British participants were presented with a Native American folk story. When asked to recall the story, the participants’ recollections omitted elements that were culturally foreign to them—they transformed the story to be consistent with their own cultural knowledge, such as changing the word ‘canoe’ to ‘boat’. This means that our prior experiences and memories shape our perception and contribute to constructing new memories. In fact, we perceive and solidify information as new memories more easily when it is similar to existing mental representations than when it diverges from pre-existing mental representations. Like schemata, scripts are also abstract structures, but ones that include an “appropriate sequence of events in a particular context” (Schank & Abelson, 1975, p. 170).14 Familiar examples of scripts include ‘riding a bike’ or ‘tying shoelaces’, which involve a well-delineated chronological sequence of events, but ones that are difficult to verbalize. Because of their chronological organization, scripts are often characterized as procedural memory. While classical conditioning guides future behavior by associating two stimuli, scripts instead guide future behavior (and predict future events) by engendering expectations that concern the appropriate or most likely sequence of actions. In contrast to implicit memory, explicit memory can be consciously and intentionally retrieved. Scholarship on explicit memory subdivides it into semantic memory (collectively shared knowledge that includes facts, concepts, and names) and episodic memory (autobiographical information only accessible by a single individual, including events and experiences). Semantic memory is what we commonly think of as knowledge that rests upon culturally shared categorizations—countries’ capitals, sports’ rules, musical genres.15 Whereas semantic memory draws on socially constructed and shared categories, episodic memories draw on personal experiences; hence, we commonly regard these as recollections—the foods we ate this morning, our first day at school. Ecological/Evolutionary Perspective

The ecological/evolutionary approach to memory has coexisted with the record-keeping, working memory, and long-term memory theories. This approach moves further away from the exploration of memory as ‘storage’ or ‘processes’ and toward an understanding of the ‘function’ of memory—instead of focusing on how memory works or where in the brain memories reside, an ecological/evolutionary perspective focuses on why we have memories. From this perspective, memory’s primary function is not to preserve past experiences but to anticipate future events (Morris, 1988), providing vital ecological and evolutionary advantages that allow us to predict future outcomes and adapt to novel environments. Because of

Appendix IV: Memory

191

this anticipation mechanism, our memory constantly adapts by registering salient information or deviations of previously consolidated patterns, reconfiguring our cognitive systems in response to new experiences. Here, the saliency or distinctiveness of stimuli plays a critical role in their memorization. Structural saliency is a byproduct of chunking and recognition. Structurally salient events are noticeable deviations from gestalt or schemata patterns; these attract attention and thus are more easily consolidated in long-term memory (Alger et al., 2018; Holmes, 1972; Waters & Leeper, 1936). Cohen and Carr (1975) explored this phenomenon by presenting photographs of human faces to participants and requesting they judge the distinctiveness of each face; in a subsequent stage of the experiment, participants more accurately recognized the faces they rated as distinctive. As Guenther (1998) succinctly notes, “making information distinctive or associating information with distinctive images and ideas can promote better memory of that information” (p. 328). Semantic and affective saliency, on the other hand, is a byproduct of activations and associations. Guenther (1998), for instance, notes that “events associated with strong emotions . . . are better remembered than emotionally more neutral events” (p. 323). Similarly, Alger et al. (2018) mention that memory of an “emotionally salient [episode] is preferentially preserved, whereas memory for neutral, contextual detail is forgotten or suppressed” (p. 34). This facility to adjust to changes in our environments equips us to predict future events, resulting in ecologically advantageous behavior. This suggests a close relationship between memory, imagery, and expectancy. Eysenck and Keane (2015), for instance, highlight the interplay between memory and imagery when noting that “imagining future events involves the same (or similar) processes to those involved in remembering past events” (p. 274). Huron and Margulis (2010), in turn, highlight a common goal of memory and expectancy in noting that the evolutionary and ecological goal of expectancy is “to prepare an organism for the future, not the past” (p. 579) and that the “adaptive purpose of memory must be prospective rather than retrospective” (p. 580).16 Therefore, like general memory, musical memory functions to aid expectation and prediction. Chapter 6 fleshes out how these ecological and evolutionary roles of memory play a role within the context of film music leitmotifs.

Notes 1. Although this theory gained traction in the 1970s with the development of computer technologies and artificial memory systems (e.g., hard drives, tapes), recent paradigm shifts in cognition and psychology have challenged the record-keeping conception of memory. 2. Note that such retrieval of sensory memories does not entail any complex processing. 3. Once sensory information reaches our brains, it can disintegrate or, if our brains register it as relevant, it can become part of the memory system that has been traditionally parsed into long- or short-term memory. 4. Although this distinction may seem arbitrary, much research in neuropsychology supports it, particularly through the study of individuals with brain damage or amnesia—individuals with amnesia generally present long-term memory impairments while maintaining a well-functioning short-term memory (Spiers et al., 2001). 5. A ‘connectionist’ model of memory (McClelland, 2000) also rejects the traditional storage view, favoring activity patterns between neuronal units that form neural networks. Memorization entails changes in our bodies and brains. Although Barlett introduced the connectionist approach to

192

6.

7.

8. 9.

10. 11.

12.

13. 14. 15.

Appendix IV: Memory

memory in 1932, much recent research from cognitive neurophysiology supports a connectionist approach by revealing that “memory reflects changes to neurons involved in perception, language, feeling, movement, and so on” (Guenther, 1998), thus outright dismissing the division into short- and long-term memory systems (Macken et al., 2003; Nairne, 2002). As a result, working memory does not consist of a specific neural substrate or dedicated brain module; instead, it consists of “distributed, overlapping and interactive cortical networks that in the aggregate encode the long-term memory” (Fuster & Bressler, 2012, p. 207). Similarly, Craik and Lockhart (1972) identified various “levels of processing” to consolidate information in longterm memory, ranging from shallow recognition to deep semantic analysis, with “deeper levels of analysis produc[ing] more elaborate, longer lasting and stronger memory traces than shallow levels” (Eysenck & Keane, 2015, p. 230). Imaging studies support the interplay between working memory and long-term memory, showing interactions between large-scale brain networks during general memory tasks (Fuster & Bressler, 2012). Real-world scenarios illustrate a close interplay of various memory systems. Kvavilashvili and Mandler (2004), for instance, explored the spontaneous (involuntary) recall of semantic memories; they report on a diary, retrospectively finding cues that triggered those semantic memories. Their findings suggest that involuntary semantic memories are triggered by our subliminal perception of related information—as an example, they mention the involuntary “mindpopping” recall of the names Itchy and Scratchy (from The Simpsons) when one of the authors scratched her back. Additionally, evidence from cognitive neuroscience suggests that working memory engages brain areas involved in long-term memory (D’Esposito, 2007). For instance, imaging studies identified simultaneous activity in brain areas critical for working memory and long-term memory encoding (Wagner et al., 1998), suggesting that processes within working memory temporarily activate longterm memory systems (Ward, 2015). Structural congruency also promotes memorization and recall. An informal experiment conducted by my graduate students illustrates this phenomenon of false memories. After a feature presentation, filmgoers were asked about the music in the film, and most responses addressed the music’s “beautiful melodies”, “nice bass”, “emotional impact”. Paradoxically, the film contained no music at all. This reflects that an experience is ‘reconstructed’ based on existing schemata: since most films contain music, filmgoers reconstruct and recall their experience shaped by schemata based on such prior knowledge. Tulving’s (1983) ‘encoding specificity hypothesis’ purports that the presence of contextual similarity surrounding the initial encoding and subsequent retrieval of memories helps consolidate them as long-term memory traces. Some scholars, however, understand explicit and implicit memory as along a continuum, with no boundaries between them (e.g., Challis et al., 1996). This line of thought suggests that explicit memory engages in deeper levels of processing, as compared to implicit memory, and that implicit memory serves as a precursor to explicit memory. Much empirical research supports this theory; for instance, using eye-tracking and measuring gaze fixation, Ryan et al. (2000) explored how participants form associations via implicit learning when exposed to images of real-world scenes; they concluded that “memory representations of scenes contain information about relations among elements of the scenes, at least some of which is not accessible to verbal report” (p. 454). In a similar vein, Juslin and Västfjäll (2008) describes ‘evaluative conditioning’ as “a special kind of classic conditioning that involves the pairing of an initially neutral conditioned stimulus (CS) with an affectively valenced, unconditioned stimulus (US). After the pairing, the CS acquires the ability to evoke the same affective state as the US in the perceiver” (p. 564). In Pavlovian conditioning, once two stimuli are associated, either stimulus in isolation may trigger the same physiological response. Scripts are also called ‘procedural memory’. Semantic memory seems to be amodal (i.e., not tied to any particular modality of perception), possibly constructed through a process of “transduction” (Barsalou, 2008), where initial modal representations turn into amodal representations. For instance, when encountering dogs, “modal representations arise as dogs are seen, heard, and touched. In turn, these modal representations are transduced into amodal symbols that stand for these experiences” (p. 92). There is, however, no definitive empirical evidence for the presence of amodal brain representations; this suggests

Appendix IV: Memory

193

that, during retrieval of semantic memories, the original modal representations may become active. It is therefore currently unclear how semantic categories are represented in the brain, and how we arrive at stable category representations. In fact, Barsalou (2008) offers support for the involvement of perceptual and motor neural substrates when engaging semantic categories. 16. Huron and Margulis (2010) regard various forms of memory (including semantic, episodic, and procedural) as “different forms of expectation” (p. 581).

APPENDIX V Auditory Perception

We are constantly submerged within a complex sonic environment that floods our auditory sense, yet we navigate this sonic environment seamlessly and with little cognitive effort. The sounds of rustling leaves, birds, rain, wind, conversations, music, traffic, may all be present at once, yet we have the capacity to distill overlapping sounds, to attend to one or more of them, or even to ignore them altogether. This appendix explores vital mechanisms of auditory perception that allow us to engage meaningfully with our sonic environments, mechanisms that pertain to Auditory Stream Analysis. Through this exploration, we will begin to recognize that auditory perception is a critical component of an evolutionarily advantageous cognitive mechanism we will call ecological resonance. Auditory Stream Analysis

Perception is automatic, subliminal, and hardwired. It entails using sensory information to form useful mental representations of our environments, mental representations that ultimately guide and influence our behavior. Bregman (1990) offers a comprehensive account of various aural perception processes and frames these within the umbrella of Auditory Stream Analysis (ASA).1 Through Bregman’s writings, we learn that our perception is not a replica of the natural world; instead, biological constraints and past experiences modulate our perception—we process an auditory landscape by relying on innate and learned mechanisms to recognize regularities, partition and group information, and ultimately generate a mental representation. ASA entails two primary and interdependent processes: (1) the bottom-up parsing of sensory information, driven by gestalt principles, and (2) the top-down matching of structures embedded in the incoming stimuli, driven by schemata and semantic units.2 Bottom-up and top-down processes both draw on schemas, but the schemas corresponding to each present marked differences.3 Bottom-up processing draws on ‘innate’ schemas— these are governed by gestalt grouping principles, allow us to attend to surface-level

Appendix V: Auditory Perception

195

regularities, and are finite and shared by all humans. Top-down processing, on the other hand, draws on ‘learned’ schemas—these are mental representations formed by repeated exposure to stimuli that present some degree of regularity, allow us to attend to abstract and conceptual-level patterning, and are potentially infinite and unique to individuals or communities. Bottom-up and top-down processing are thus functionally distinct, the former geared toward constructing perceptual entities, the latter toward selecting information by activating learned schemas. Although seemingly independent, bottom-up and top-down processing work in tandem during auditory perception to enable us to establish (and later derive) meaning from a sonic landscape. Figure V.1, for instance, offers a visual analogy that distills the bottom-up and top-down processing that allows us to go from perception to the generation of an action plan.4 First, the interaction of top-down and bottom-up processing prompts us to identify a string of letters on the left box; at this stage, however, this string of letters does not carry any meaning. Then, with the support of innate gestalt principles, we parse out certain letters belonging together because of their unique font. Subsequently, based on learned schemas, we recognize two meaningful words. Ultimately, we draw on these words’ semantic content to derive a useful message, one that prompts us to generate an action plan. In spoken language, innate schemas allow us to discriminate amongst surface-level regularities to identify speech sounds. Learned schemas, on the other hand, account for deep- and abstract-level regularities—from the phonetic (the sound for a single letter) to the syntactical (sentence structure), the stylistic (colloquial, poetic), and even the narrative structure (linear narration, circular stories). Analogously, in music perception, innate schemas attend to the surface-level regularities that allow us to identify sounds as music, and learned schemas account for deep- and abstract-level regularities—from the phonetic (sound of a violin, sound of a marimba) to the syntactical (harmonic progression, formal design), and even to the stylistic (pop, classical). Because gestalt theory attends to the principles that govern sensory perception, it is especially useful for conceptualizing bottom-up mechanisms. Gestalt psychology maintains that our tendency to construct patterns is automatic, pre-attentive, and grounded in a basic set of hardwired schemas that conform to fundamental grouping principles. These principles, which include proximity, similarity, closure, good continuation, common fate, and good form, allow us to process ‘chunks’ of information or patterns. In music, these grouping principles operate on our perception of both contiguous and simultaneous sounds—broadly understood as the music’s melodic and harmonic dimensions, respectively—and emerge from multiple musical and sonic parameters, including pitch (frequency), rhythm (temporal proximity), timbre and articulation (spectral information), dynamics (loudness), and source (positioning in the sonic landscape).

FIGURE V.1

From perception to action.

196

Appendix V: Auditory Perception

Proximity

The gestalt principle of proximity accounts for our tendency to group objects that feature relative nearness in space, time, or another parameter. Visually, we will perceive the dots on the left of Figure V.2 as randomly scattered, while we will perceive the same number of dots on the right as grouped into two distinct units because of their spatial proximity.

FIGURE V.2

Gestalt principle of proximity.

In music, proximity may manifest itself through the music’s various parameters—e.g., sounds within a particular frequency range, sounds in close temporal adjacency, sounds emanating from a relatively defined location. For instance, by exploiting the principle of proximity in range, composers may create the impression of multiple melodies emanating from a single line that alternates between two relatively distant registers—musicians call this a ‘compound melody’.5 Similarity

The gestalt principle of similarity accounts for our human tendency to form groups based on resemblance. When observing the shapes in Figure V.3, the gestalt principle of proximity will prompt us to perceive three distinct groups; within each of these groups, however, our brains will subconsciously form two groups based on the shape’s similarities in terms of their size (left), contour (center), or color (right). In music, multiple parameters may contribute to forming groups based on similarity, including timbre, loudness, and articulation. For instance, in much orchestral music, composers create the impression of a continuous, single, unified musical gesture via instrumental ‘dovetailing’—superimposing the endings and beginnings of segments performed by

FIGURE V.3

Gestalt principle of similarity.

Appendix V: Auditory Perception

197

instruments with a similar timbre to render musical figures that extend through a range that falls outside of any single instrument. Common Fate

The gestalt principle of common fate is related to the principle of similarity, but instead of forming groups based on static attributes such as shape, size, and color, we group elements based on dynamic attributes such as spatial behavior, including movement, direction, and onset. In Figure V.4, the lines depict the spatial behavior of four elements, prompting us to group them as a single unit (left), as two distinct units (center), or as four distinct units (right).

FIGURE V.4

Gestalt principle of common fate.

In music, our perception of groups based on the principle of common fate is influenced by the contours, durations, and onsets of various musical layers. The common fate principle thus allows for our perception and recognition of heterophonic, homophonic, and polyphonic musical textures: heterophonic texture will result from all layers sharing analogous onsets, durations, and contours; homophonic will result from our perceiving two layers, one functioning as melody, the other as accompaniment; and polyphonic will result from perceiving multiple independent layers that do not share analogous onsets, durations, or contours. Masking and Good Continuation

The gestalt principles of masking and good continuation account for our tendency to assume that an element continues underneath another element that perceptually masks it. When looking at Figure V.5 on the left, we will assume the horizontal rectangle continues behind the triangle, even if a portion of the rectangle is perceptually inaccessible; and, on the right, we will assume the triangle is complete and behind the rectangle, even if part of the triangle is perceptually inaccessible.

FIGURE V.5

Gestalt principle of masking and good continuation.

198

Appendix V: Auditory Perception

The perception of masking and good continuation in music is contingent (primarily) upon the loudness of various layers. For instance, a loud tam-tam hit will temporarily mask a sustained pianissimo sonority in the strings—when masked by the louder sounds of the tamtam, the pianissimo’s sustained sonority is inaccessible to our perception, yet we assume its continuity.6 Some composers use this principle to give the impression of a continuous layer, particularly when arranging orchestral music for piano. Figure-Ground

The gestalt principle of figure-ground grouping prompts us to engage our attention selectively. In the famous profiles/vase figure, we will attend selectively to either the profiles or the vase, but not to both simultaneously. (See Figure V.6.) In auditory perception, this gestalt principle is most evident in the phenomenon known as the ‘cocktail party effect’—even when embedded within a complex and dense acoustic landscape, such as in a cocktail party, we can selectively direct our attention to a single stimulus (e.g., a conversation or the music) and regard other aural stimuli as background. For instance, in homophonic music, we perceive a melody as the figure and the accompaniment as the background. In music perception, this principle is most evident when focusing our attention on different parts of a texture—in polyphonic music, for instance, we may consider any voice or part as figure or (back)ground; yet, while we can attend to a single melodic strand with relative ease, attending to multiple simultaneous strands will present a cognitive challenge.7

FIGURE V.6

Gestalt principle of figure-ground.

Saliency

Although not a gestalt principle of perception, saliency interacts with innate and learned schemas. In fact, saliency results from our “processing of difference in the context of similarity” (Hunt, 2013, p. 10), from perceiving deviations from a trend or norm both within

Appendix V: Auditory Perception

FIGURE V.7

199

Saliency in bottom-up and top-down perception.

bottom-up and top-down perception. In Figure V.7, when looking at the shape on the left, our attention will be drawn to the small lump as a deviation from a smooth trend; and, when reading the text on the right, attention will be drawn to the word “sexophone”, which deviates from the expected ‘saxophone’ by embedding the word ‘sex’ and hence introducing sensual resonances. Because we are most efficient in processing information that conforms to innate or learned schemas, information that features some degree of saliency or distinctiveness will add to the cognitive load when processing such information.8 In music, composers may draw on our tendency to react to bottom-up saliency when deviating from an established pattern—a corner in a melodic contour, an unusual chord in a repeating harmonic pattern, a syncopation in an otherwise non-syncopated metrical and rhythmic framework.9 In turn, performers may draw on our tendency to react to top-down saliency when deviating from a learned schema, such as when re-harmonizing a standard or when inserting agogic accents for expressive purposes.10 Ecological Resonance, Integrating Schemas

Innate and learned schemas interact during perception. The foundational presence of innate schemas is a necessary precondition for the emergence of learned schemas. The interaction between innate and learned schemas thus provides us with an optimal strategy for adapting to the environment because: A system that did not take into account the sensory input [i.e., did not engage bottom-up mechanisms] would be cut off from the outside world, whereas one that did not use previously acquired knowledge [i.e., did not engage top-down mechanisms] would have a very unstable representation of the changing world. (Bey & McAdams, 2002, p. 852)11 Such interactions between innate and learned schemas permeate our evolutionary history, working below our conscious attention, allowing us to navigate the sonic environment, where sounds are not mere patterns but meaningful cues—this interaction mechanism is best described as ecological resonance. Bregman (1990) addresses our tendency to ecologically resonate with everyday sounds, noting that “our brains are not built to hear sounds in the abstract, but to form descriptions of environmental events” (p. 679). Similarly, in music, Reybrouck (2010) anticipates the notion of ecological resonance, recognizing the interaction between innate and learned schemas in “the construction of new distinctions and observables as well as the recognition of knowledge structures that are already acquired as the outcome of previous interactions [with sounds]” (p. 189).12 This interdependency between bottom-up and top-down processes, as highlighted by Bregman and Reybrouck, brings us nearer to understanding auditory perception’s evolutionary and ecological purposes.

200

Appendix V: Auditory Perception

Gibson (1966) addresses bottom-up and top-down perception as hardwired reflexes in the visual domain that bestow organisms with an ecological advantage.13 In essence, this mechanism entails structuring information into objects we recognize. Please turn to Figure V.8 before reading further. Although this image may initially appear to be just the surface of a tree bark, we will identify the frog in it upon engaging our top-down schemas. Furthermore, once we recognize the frog, we cannot ‘unsee’ it, and our attention will always be drawn to it. Although animals try to take advantage of their camouflaging mechanisms, they also gain an equal ecological advantage by perceiving a predator or prey to enable appropriate adaptive behavior. In terms of everyday sounds, for instance, a mosquito’s buzz is not just meaningful because we have formed a schema upon repeated encounters with the sound, but because we associate the schema with the presence of a mosquito and the potential for a mosquito bite. As a result, an organism that associates a sonic learned schema with an environmentally threatening situation may initiate adaptive behavior when encountering such sound. Reybrouck (2010) argues that music functions as environmental sound; therefore, just as organisms adapt in response to their sonic environments, we “show adaptive behavior in [our] interactions with the music as environment” (p. 189). Likewise, Biancorosso (2010) traces a mechanism suggestive of ecological resonance to our evolutionary history and mentions that sound can signal the unknown or ominous, such as a distant menace or a fast-approaching threat requiring immediate response . . . insofar as the music partakes of the creation of such scenarios, it may be placed alongside memories of sounds heard in real-life situations and thus be treated, at least indirectly, as a natural sign. Our response to such a sound would have to be, at least to some extent, wired into the brain. (p. 317) Chapter 6 expands on our hardwired tendency to ecologically resonate with a film’s sonic and musical landscape.

FIGURE V.8

Ecological function of perception.

Appendix V: Auditory Perception

201

Notes 1. Bregman (1990) labels his theory “Auditory Scene Analysis”. My changing ‘scene’ for ‘stream’ is intended to avoid confusion with the use of the word ‘scene’ in the context of a film. Bregman’s theories draw on Gibson’s (1966) ecological theories of visual perception and Gregory’s (1980) top-down processing theory. 2. Bottom-up auditory perception begins with the stimulus itself; it is a data-driven process carried in one direction, from auditory organs to the auditory cortex. Top-down auditory perception, on the other hand, uses mental images and contextual information to process auditory stimuli. Bregman (1990) addresses bottom-up and top-down processes respectively as ‘primitive’ and ‘schema-based’: “It seems reasonable to believe that the process of auditory scene analysis must be governed by both innate and learned constraints . . . the effects of the unlearned constraints will be called ‘primitive segregation’ and those of the learned ones will be called ‘schema-based segregation’ ” (p. 242). Regardless of their labeling, all scholars agree that these interacting mechanisms are hardwired and pre-attentive, meaning that we are born with the capacity to engage them and that we do so below the level of consciousness. 3. Schemas are mental structures that emerge from regularities we perceive in our environment. Within the context of this book, I define schemas quite broadly to include hard-wired conceptual frameworks and gestalt principles of organization. 4. Since illustrating mechanisms of aural perception within a book is best achieved by drawing on analogies with visual perception, all explanations in this appendix include a visual component before considering its aural counterpart. 5. Bregman (1990) surveys various experiments and concludes that “if the alternation is fast enough and the frequency separation great enough, listeners will not experience a single stream of tones alternating in pitch, but will perceive two streams of tone . . . When two streams are heard, the listener has the impression of two different sources of sound” (p. 642). 6. Bregman (1990) points out that, even if a sustained sound were to be “deleted and replaced by a much louder sound, the actual existence of the background pattern is irrelevant to our perception of good continuation” (p. 346). 7. Drawing selective attention to the figure does not mean, however, that the background has no effect on our perception. On the contrary, the background greatly influences our perception of the figure, particularly in music. 8. At a neural level, experiments using event-related potentials (ERP) illustrate that our neurons fire differently when exposed to a stimulus that deviates from a pattern; for example, even a small deviation of a metronomic pulse will trigger a mismatch negativity response that can be observed in the electroencephalography (EEG) signal (Zanto et al., 2006). 9. Dyson and Watkins (1984) conclude that “corners of melodic contours act as features and are perceptually more salient than the intervening notes” (p. 483). 10. In contrast to innate schemas, learned schemas “[do] not perceptually segregate sounds, creating perceptual units such as auditory streams, but allow us to select information from a mixture by a matching process between schemas stored in memory and a sensory representation” (Bey & McAdams, 2002, p. 851). Observing saliency through learned schemas, therefore, results not from what we hear, but instead from a discrepancy between what we hear and what we expect to hear. 11. Bey and McAdams (2002) identify both the ‘attentive’ nature of learned schemas, “when we explicitly try to hear a sound source or a sound sequence in a background mixture” (p. 845), as well as their ‘pre-attentive’ nature, when “being in a room with many people talking and hearing one’s name emerge from the mixture” (p. 845). In addition, empirical studies conducted by Boltz (2001) show that schemas lessen the cognitive load by “reduce[ing] the amount of attentional effort that is needed for perception and comprehension, and organize[ing] one’s perceptual experience into a coherent and intelligible whole” (p. 432). While Bregman (1990) claims that innate and learned mechanisms are different because the former is “automatic” and the latter involves “attention”, I argue here that both are automatic. 12. Reybrouck (2010) further highlights the ‘online’ confluence of the perceptual immediacy of gestalt-based innate schemas and the necessary conceptual abstraction of learned schemas, noting the “dynamic tension between ‘experience’ and ‘recognition’ with the former relying on a moment-to-moment scanning of sensory particulars, and the latter relying on processes of abstraction and generalization” (pp. 187–88).

202

Appendix V: Auditory Perception

13. Visual and auditory perception rely on analogous grouping strategies to minimize the cognitive and attention loads and to help us cope with vast amounts of information reaching our senses. Furthermore, our human tendency to integrate visual and aural features appears to be innate (or at least prelinguistic). Wagner et al. (1981) found that infants develop visual and auditory integration before learning to talk; in their experiments, infants attended to dotted lines when exposed to a series of short sounds and to solid lines when exposed to long sounds. In a study that extends Marshall and Cohen’s (1988) seminal work and revisits their Congruence-Association Model, Millet et al. (2021) explore this phenomenon within film music with the use of eye-tracking technology; they explain that, during moments of structural congruency, the “music led to quicker first fixations on film objects . . . [accentuating] the saliency of film objects and [supplying] emotional information altering the viewers’ sentiment towards the film” (p. 1).

APPENDIX VI Archetypes

We are surrounded by archetypes, universal symbols that emerge from within a culture and become units of meaning shared and reproduced by its members. In literary works, gardens usually symbolize love and fertility, towers symbolize power and worship. In films, shadows often symbolize dark forces or malicious tendencies, bridges symbolize transitions between opposites. Music partakes in this universe of archetypes with surface-level figures, timbres, and even styles, expressing our cultures’ settings, rituals, and structures.1 Within musicosemiotic scholarship, these culturally wrought units of meaning are called musical ‘topics’.2 This appendix traces the origins of topics theory and surveys its application to a wide range of repertoires, from eighteenth-century concert music to twenty-first-century popular and film music. It then explores musical topics through the lens of modern semiotics, whose frameworks provide an elegant inference engine that allows us to glean the potential mechanisms whereby we derive meaning from topics. Although such frameworks rest solely on speculative insights, they have propelled a wealth of empirical research on musical topics. Therefore, this appendix surveys relevant empirical research to expand our awareness of musical topics’ expressive power. Topics Theory

Composers and scholars have been keen to identify musical units that suggest extra-musical meaning.3 Early music treatises (e.g., Heinichen, 1711; Mattheson, 1739) discuss compositional practices using ‘loci topici’—surface-level musical textures and gestures with rhetorical and affective potential.4 Ratner (1980) drew on these treatises to compile a thesaurus of characteristic figures encultured listeners of the time would recognize, laying the groundwork for numerous scholars to refine and apply topics theory to various repertoires. Agawu (1991) distills a compendium of the most frequently used topics in eighteenth-century Western concert music, a ‘Universe of Topics’.5 The topics in this compendium reflect a broad range of settings, which we can further categorize as affective states, musical figures and gestures, styles and genres, and social contexts

204

Appendix VI: Archetypes

FIGURE VI.1

Agawu’s “Universe of Topics”. (After Agawu, 1991)

and functions. Figure VI.1 presents various topics from Agawu’s compendium arranged within representative categories. For most eighteenth-century listeners, the topics in Agawu’s compendium would have been readily recognizable; but in the twenty-first century, because we are distant from the musical conventions of the time, we need “to be schooled in the idiom of the eighteenth century” to develop the necessary “listener-competence” (Agawu, 1991, p. 49).6 An encultured listener, then and now, would recognize surface-level features in Mozart’s Piano Sonata K. 332 and hear a succession of musical topics rather than merely a series of musical themes or figures. Figure VI.2 presents a portion of this work and depicts the topics suggested by the embedded musical figures and gestures. The pace and range of the melody in bars 1–4 suggest the ‘singing style’ topic, while the triple meter and the accompaniment featuring a tonic pedal tone suggest the peaceful ‘pastoral’ topic. The emboldened, mezzo-forte contrapuntal figures in bars 5–8, which contrast with the homophonic texture of the preceding bars, suggest a sudden shift to the ‘learned style’ topic; yet toward the end of this figure, two staccato chords seem to poke fun of the prior serious tone, suggesting a possible ‘opera buffa’ topic. Emulating the sound of hunting horns, featuring dotted rhythms and the characteristic 3rd-5th-6th intervallic gesture in a major mode, bars 12–22 suggest the ‘horn call’ and ‘hunt style’ topics. Then, at bar 22, the sudden forte dynamics, the shift to the relative minor mode, a more active rhythmic profile, and the diminished-seventh sonorities suggest the turmoil and instability characteristic of the ‘Sturm und Drang’ [‘storm and stress’] topic.7 Resembling Agawu’s compendium for the common practice repertoire, Tagg’s (1999) compendium of “feels”, which he defines as an “ethnocentric selection of possible connotative spheres” (p. 12), includes topics listeners would recognize in more popular styles.8 Just like an eighteenth-century audience member would subliminally recognize the ‘Sturm und Drang’ or ‘hunt style’ topics in concert music, nearly every twenty-first-century audience member would recognize the ‘Spaghetti Western’, ‘twinkling happy Christmas’, or most other topics included in Tagg’s compendium. Figure VI.3 offers a selection of Tagg’s extensive list of feels.

Appendix VI: Archetypes

FIGURE VI.2

205

Mozart’s Piano Sonata K. 332, measures 1–26.

Because topics are conventionalized signs, “competence is assumed on the part of the listener, enabling the composer to enter into a contract with [the] audience” (Agawu, 1991, p. 33).9 Film composers, for instance, are exceptionally attuned to twenty-first-century topics, particularly because most film directors communicate with their film composers by employing words or phrases similar to those included in Tagg’s compendium; composers, in turn, must recognize and translate these into music that evokes the desired response from the audience.

FIGURE VI.3

Tagg’s “ethnocentric selection of possible connotative spheres”. (After Tagg, 1999)

206

Appendix VI: Archetypes

Within the film music repertoire, such conventionalization of topics began during the silent film era’s film-accompaniment practice. From the mid-1890s to the late 1920s, several music anthologies or encyclopedias circulated among film musicians, primarily accompanists. These volumes contain music classified according to ‘moods’ or ‘dramatic settings’ and include several examples in each category from which accompanists would select to support the narrative. Although to our modern ears we may hardly conceive of some of the included categories as topics (such as ‘Alabama’, ‘Bees’, or ‘Chatter’), from a contemporary listener’s perspective, all these categories conformed to a comprehensive universe of topics. Ernö Rapée’s (1925) Encyclopedia of Music for Motion Pictures outlines fifty-two categories, capturing the most frequently used dramatic settings in films (at the time). In the introduction to the volume, Rapée notes: “In creating fifty-two divisions and classifications in this Manual, I tried to give the most numbers to those classes of music which are most frequently called upon to synchronize actions on the screen”. (iii) The categories are varied, with “One-third of all film footage [used] to depict action; another third will show no physical action, but will have, as a preponderance, psychologic [sic] situations; the remaining third will neither show action nor suggest psychological situations, but will restrict itself to showing or creating atmosphere or scenery”. (iii) Rapée also offers suggestions for the suitable placement of these selections. The music in the ‘sinister’ category, for instance, “is meant for situations like the presence of the captured enemy, demolishing of a hostile aëroplane or battleship, or for the picturing of anything unsympathetic” (iii); the selections under ‘parties’ may be “suitable also for the portrayal of social gatherings in gardens” (iii). Most music anthologies for motion pictures of the early twentieth century drew almost exclusively from pre-existing Romantic and post-Romantic concert music. This practice perpetuated and reinforced already-established concert-music topics, introducing them to the film music repertoire.10 As a result, these topics, and the associated surface-level figures, became profoundly influential to film composers of the immediate decades that followed, with film composers continuing to emulate these topics during the embryonic stages of sound film. As a result, these encyclopedias directly established a set of musical topics that became inextricably linked to particular settings or film genres. Semiotic Perspectives on Topics

Anglo-American semiotics (e.g., Charles Sanders Peirce, John Searle, George Lakoff, Deborah Tannen, Mark Johnson) views meaning as a product of the relationship between the sign and the interpretant, in which the sign represents the object to the interpretant. On the other hand, continental semiotics (e.g., Ferdinand de Saussure, Roland Barthes, Jean-Francois Lyotard, Jacques Derrida) is heavily influenced by structuralism, focusing on analyzing the underlying structures and systems that shape the meaning of cultural texts. The semiotic

Appendix VI: Archetypes

207

exploration of topics can be framed within these semiotic traditions, each illuminating different facets of topics as signs. While a Peircean perspective would shed light on the epistemology of topics by situating them within a sign typology, a Barthesian perspective would shed light on the process whereby topics emerge as units of meaning within the musical discourse. Peirce’s Trivium

Peirce established three categories of signs, defined by the relationship between a sign and what it represents: icons, indexes, and symbols. Icons physically resemble what they represent—e.g., a bicycle drawing to suggest their presence on the road. Indexes point to or show evidence of what they represent—e.g., smoke to suggest there is, or has been, a fire. Symbols emerge as arbitrarily established conventions—e.g., a red traffic light to mean drivers should stop. Within Peirce’s taxonomy, topics are symbols because their signifying power rests on an arbitrarily established and conventionalized relationship between a sign and its meaning.11 Their pre-culturized meaning is particularly evident when suggesting social contexts and functions, such as in the ‘Italian’, ‘Turkish’, ‘ecclesiastical’, and ‘military’ topics. Their symbolic nature notwithstanding, topics often develop out of icons or indices, acquiring conventionality through time. For example, while resembling the acoustic rhythmic profile of galloping horses, the ‘gallop’ topic has been further reinforced within culturally defined repertoires by programmatic elements (e.g., the music’s title or lyrics). Fortunately, Peirce’s trivium allows for signs to reflect more than one type of relationship; hence, besides their primary typology as a symbol, topics can, at some level, also reflect a secondary typology. Barthes’s Structuralist Framework

Most tenets of structural semiotics draw on binary oppositions. One such opposition relates to two complementary facets of a sign: the signifier (its material form) and the signified (its mental concept). Musical topics exhibit this emblematic dimension of signs, the signifier-signified relationship—a Theremin (signifier) represents science fiction or supernatural events (signified), a contrapuntal figure (signifier) represents a learned setting (signified). However, a closer inspection reveals that they rely on a richer semiotic process, a chain of signification that can be unpacked by drawing on Barthes’s denotation-connotation framework. While a topic’s denotative meaning(s) represents its most descriptive level, its connotative meaning(s) represents its associated social overtones or cultural implications. Beneath the surface-level meaning of a musical sign (its denotative meaning), there are numerous additional layers of associative meanings (its connotative meanings). Superimposing Barthes’s denotation/connotation binary onto the exploration of topics illustrates a chain of significations, one in which a signifier is itself a sign with its own signifier—the surface musical phenomena first denote a particular topic, and in turn, the resulting construct, a new sign, connotes extra-musical and socio-cultural associations. For example, surface-level musical features (signifiers) such as minor mode, slow tempo, duple meter, repetitive quarter-note rhythm with interspersed dotted figures, and low register denote the ‘funeral march’ topic (signified); in turn, that entire topic becomes a new sign (signifier), connoting somber associations of death, mourning, cemeteries, and alike (signified).

208

Appendix VI: Archetypes

FIGURE VI.4

Chain of signification in the ‘funeral march’ topic.

Topical Markers

Within a topic’s chain of signification, the primary link rests on the listeners’ attunement to surface-level musical figures, or topical ‘markers.’ These markers help: identify a particular musical style and often, by connotative extension, the cultural genre to which that musical style belongs . . . different combinations of different aspects of duration, rhythm, timbre, tonality, spatiality, diataxis . . . [help listeners] instantaneously know if they’re hearing 1970s disco rather than zouk, rococo chamber music rather than death metal, glitch dub rather than Gregorian plainchant, mbaqanga rather than Muzak, an Elizabethan madrigal rather than a low-church hymn, a ra-ga performance rather than a romantic pop ballad, a national anthem rather than a TV detective theme. (Tagg, 2012, p. 522)12 As the primary elements within the chain of signification, markers are thus vital for encultured listeners to recognize musical topics effectively. When connoting a style or genre, a topic can share an unlimited number of markers with the style or genre from which it derives. Although this may seem to blur the lines between an identity relationship and a semiotic one, contextual clues catalyze our recognition of topics as signs. For example, a segment of film music might incorporate all markers of the ‘70s Disco Music’ topic, and thus it might seem to become a disco hit on its own. However, contextual clues help us recognize it as a topic rather than as a disco hit; these contextual clues may be internal to the film (e.g., narrative, visuals, or dialogue), external to the film (e.g., the fact that we are in a movie theater and not in a discotheque in the ‘70s), or they can be both. Although a single marker may connote a particular topic, in most instances, a combination of several markers is necessary to connote a topic with precision. For instance, the minor mode marker is insufficient to suggest a specific topic; instead, only when combined with other markers such as slow tempo, duple meter, and dotted figures, will it effectively connote the ‘funeral march’ topic. Therefore, although there is a relatively finite number of markers, they allow for an infinitely larger number of topics when used in combination.

Appendix VI: Archetypes

209

For instance, the triple meter marker may combine with other markers such as pedal tones, harmonic progressions, tempo, rhythmic figures, and phrase structures to connote an array of classical music topics, including ‘musette’, ‘minuet’, or ‘saraband.’ Empirical Approaches to Musical Topics

While the study of topics within historical musicology draws on research paradigms from the humanities to arrive at “a historical reconstruction of an interpretive competency” (Hatten, 1994, p. 3), empirical musicology draws on research paradigms from the social sciences, applying statistical and modeling methods, to gain a deeper understanding of music-related and listener-related phenomena. A subset of these studies examines the role of topics within multimedia, bringing in questions about topics’ role in establishing and advancing narrative and culture-specific themes, particularly within film. Most research converges on the idea that we each build an individualized compendium of topics and that we do so subconsciously via statistical learning. Wingstedt et al. (2008), for instance, evaluated young adolescents’ acquaintance with film music topics. In the study, participants manipulated various music parameters to adjust the music’s fit to three film scenes with contrasting narratives. Their results show that, as a part of enculturation and at an early age, we systematically and subconsciously gain an understanding of schematic musical meanings—through the repeated pairing of archetypical musical figures with specific extra-musical stimuli, we learn “through the ‘multimodal texts’ . . . from simply using media in informal situations” (p. 212). Also within film music, Bullerjahn and Güldenring (1994) investigates the effect of the ‘thriller’ and ‘melodrama’ topics in defining the listeners’ attitudes toward “the relationships between the characters and about the history and outcome of the events depicted by the film excerpt” (p. 99). They recognize that the listener’s attitudes to priorly constructed cognitive schemas related to musical topics influence their judgment about the relationships between characters. They also identify these cognitive schemas as activated by all film elements and note that: film music, especially typical genre music, produces schemas which act with other visually induced schemas (typical plot) to produce the most plausible interpretation. If the pictures are ambiguous or indefinite, the music takes on more importance in interpreting the film. (p. 102) There is, however, no conclusive evidence of how we store such statistically constructed meanings. Most theories in the cognitive sciences assume that semantic knowledge (e.g., memories, categories, symbols) resides in amodal systems, unbound from its original modality-specific representation. Musical topics, under this assumption, reside outside of the brain’s modal systems. However, this assumption entails a process of ‘transduction’ in which modal information becomes clustered and stored as amodal information. Barsalou (2008) hypothesizes a potential transduction process, arguing that, as we experience the world, “modal representations in the brain’s systems for perception, action, and affect become active . . . In turn, the brain transduces these modal representations into amodal representations that represent category knowledge in a modular semantic memory” (p. 92).

210

Appendix VI: Archetypes

FIGURE VI.5

Gendered associations of musical markers. (After Tagg, 1989)

For instance, when listening to a ‘funeral march’ topic, modal representations related to all senses arise, from the kinesthetic embodiment of the slow duple meter to the dark colors of a procession and the negative affect triggered by the circumstances. In turn, modal representations are clustered and ‘transduced’ into amodal representations associated with our experience of the music. Although such a transduction process would arguably be uniform to all humans, subtle differences in our individual experiences with topics result in a wide range of extra-musical connotations. For instance, Tagg and Clarida (2003) gathered hundreds of responses to ten brief main title themes from film and TV, seeking to discern a link between musical structures and listener connotations. They presented musical excerpts without accompanying visuals and found that music mediates wide-ranging extra-musical notions of “gender, love, loneliness, injustice, nostalgia, sadness, exoticism, nature, crime, normality, urgency, fashion, fun, [and] the military” (p. 17).13 Some studies seek to identify the source of such wide-ranging differences in connotations.14 Tagg (1989) discerns specific topical markers corresponding to collectively shared, gendered associations, presented in Figure VI.5, and reflects on the power of music to advance and construct stereotypes, warning us that “we need to know how music can make us think and feel about different sorts of people” (p. 17). Regardless of individual differences in the associations, we recognize most topical markers within seconds, or even faster. For instance, Gjerdingen and Perrott (2008) suggest that when listeners scan a radio dial, they recognize styles and genres in about a quarter of a second.15 This rapid recognition of markers contributes to the expressive power of topics to define filmic elements, such as genre, narrative setting, or character stereotypes, elements that would take significantly more time to develop via other cinematic means. Notes 1. We subliminally learn to recognize the surface-level markers (signifier) that define these musical units of meaning (signified) through consistent exposure. 2. Whereas musical topics (also known as “topoi”) are specific musical gestures with extra-musical meaning, musical archetypes may include elements such as musical forms, instrumentation, or performance techniques, which may be used ‘topically’ to convey meaning. Due to the ambiguous boundary between these two terms, within the context of this book, I use the more conventionalized “musical topic”.

Appendix VI: Archetypes

211

3. This section is not intended to summarize the state of the art of topics theory; instead, it offers the background that allows us to recognize how this analytical and interpretational framework can seamlessly be extended to explore the film music repertoire. 4. Most eighteenth-century treatises discuss topics as compositional devices and not as ‘signs’ within the musical discourse. 5. Gjerdingen (1986) identifies several musical ‘schemata’ in eighteenth-century classical music that manifest themselves as two-voice (melody-bass) paradigms. Although these schemata may elicit extra-musical associations in some listeners, thus becoming topics in themselves, Gjerdingen primarily addresses their normative use as stylistic constructs and as episodic markers within a formal structure. 6. Some studies lie at the boundary between historical and empirical musicology. For instance, Krumhansl (1998) observes that the function of topics differs significantly between listeners, even within the common-practice repertoire, and notes that while “the topics in the Mozart piece appear to function as a way of establishing the musical form . . . the topics in the Beethoven piece are more strongly associated with emotional content” (p. 119). Nevertheless, the methodological framework grounds this study within cognitive musicology, as it accounts for the experiences of present-day listeners instead of drawing on writings of contemporary composers and critics to discern the function of musical topics. 7. Topical investigation extends beyond the mere labeling of topics and explores their interaction with other musico-structural parameters. 8. Tagg refers to such musical topics as ‘paramusical symbols’ in other publications. 9. Other studies tangentially touch upon the potential of music to mediate socio-cultural conventions (Lastinger, 2011; Shevy, 2008), broadly suggesting that “the more typical the music stands for a certain film or (pop) music genre, the more clearly it triggers stereotypes and relatively sharp supra-individual expectations” (Herget, 2021). From a quasi-experimental perspective, Clynes (1977) proposes that music may embed “sentic types”, which he recognizes as surface musical features associated with particular emotions; although scentic types seem to align seamlessly with musical topics, they work at an embodied level, eliciting emotions in the listener, and thus are removed from the culturally established associations characteristic of topics. 10. Most anthologies, however, were limited in the function of the musical accompaniment and did not include music that may serve as episodic markers (opening, closing, connecting scenes) or style quotations (classical, jazz, folk, etc.). Tagg (1989) offers a reduced number of categories representative of most anthologies, which include “ ‘animals,’ ‘bright,’ ‘bucolic,’ ‘children,’ ‘comedy,’ ‘danger,’ ‘disaster,’ ‘eerie,’ ‘exotic,’ ‘fashion,’ ‘foreign,’ ‘futuristic,’ ‘grandiose,’ ‘happy,’ ‘heavy industry,’ ‘humour,’ ‘impressive,’ ‘light action,’ ‘melancholy,’ ‘mysterious,’ ‘national,’ ‘nature,’ ‘open air,’ ‘panoramic,’ ‘pastoral,’ ‘period,’ ‘prestigious,’ ‘religious,’ ‘romance,’ ‘sad,’ ‘scenic,’ ‘sea,’ ‘serious,’ ‘solitude,’ ‘space’ (cosmos), ‘sport,’ ‘suspense,’ ‘tenderness,’ ‘tension,’ ‘tragic,’ ‘travel,’ ‘water’ and ‘western’ ” (p. 24). 11. There is passionate disagreement among Peircean scholars regarding the true nature of signs. This disagreement stems, in part, from Peirce’s embarking on a meticulous classification that includes sixty-six types of signs. 12. Tagg (2012) defines diataxis as “narrative form patterns” such as twelve-bar blues or sonata form, which may enjoy a connotative dimension within interpretative communities. Similarly, Huron (2006) address formal structure, phrase construction, and cadences as “style-distinguishing” patterns. 13. In a similar study, Tagg (1989) observed a statistically significant clustering of responses gathered within cohesive semantic fields broadly defined in terms of categorizations stemming from canned-music libraries and anthologies for silent film accompaniment. 14. Wingstedt et al. (2008) suggest that, to a great extent, these differences stem from musical training and from (possibly) gendered “differences in habits of listening to music, watching movies and playing computer games” (p. 211). 15. This leads Huron (2006) to conclude that “long-term structure is not a promising creative tool. New genres or styles are better established by employing distinctive timbres, textures, or momentto-moment note transitions” (p. 208). Therefore, composers seeking to construct effective topics should attend primarily to timbre and texture, rather than harmonic progressions or rhythmic figures.

APPENDIX VII Conceptual Integration

When hypothesizing, conjecturing, imagining, envisioning, deducing, and even when reasoning, we subconsciously blend multiple mental images to construct probable scenarios, generating novel and highly subjective interpretations. Fauconnier and Turner’s (2002) Conceptual Integration theory provides an elegant model that sheds light on how these cognitive mechanisms unfold in our minds.1 This appendix reviews the foundations of conceptual blending, considers relevant neuroscientific research that supports this theory, and presents a cursory overview of music-related scholarship drawing on conceptual blending to support hermeneutic interpretations. Modeling Conceptual Blends

To understand how one creates mental representations, Fauconnier and Turner (2002) developed a model wherein mental spaces dynamically capture information as our thoughts unfold, propelling further projections and mappings among them.2 For example, when we think of a ‘sports car’, we are blending the mental spaces of ‘sports’ and ‘car’ to create a new mental space that contains elements from both. This blend allows us to understand the concept of a sports car and how it is different from a regular car. Within their model, at least two input spaces capture information from various stimuli; a generic space reduces that information into an abstraction; and a blended space superimposes the information from input spaces onto the abstract thought in the generic space.3 Fauconnier and Turner illustrate the workings of this model through the analysis of a riddle: A Buddhist Monk begins at dawn one day walking up a mountain, reaches the top at sunset, meditates at the top for several days until one dawn when he begins to walk back to the foot of the mountain, which he reaches at sunset. Make no assumptions about his starting or stopping or about his pace during the trips. Riddle: Is there a place on the path that the Monk occupies at the same hour of the day on the two separate journeys?4

Appendix VII: Conceptual Integration

FIGURE VII.1

213

Blended space to solve the riddle of the monk. (After Fauconnier & Turner, 2002)

The riddle of the Monk is not a metaphorical statement; instead, it prompts us to construct a hypothetical structure that is impossible to materialize: a mental representation in which the monk is traveling in both directions on the same day. In such a hypothetical scenario, monk one (traveling in one direction) meets monk two (traveling in the opposite direction)— the time and place of this event answer the riddle. Figure VII.1 illustrates the thought process that solves the riddle.5 Within this four-space model, elements within each space have their counterpart in another space. These elements are represented iconically or listed in an abbreviated form inside the corresponding input space: the direction on the journey, the day of the journey, the mountain with its peak and foot, and the monk on the mountain. In this conceptual blend, the generic space posits the abstract structure that relates the elements of the input spaces: a path, a moving individual, a position along the path, a day of travel, and an unspecified travel direction. Each element in one input space maps onto a corresponding element in the other input space. The blended space superimposes both input spaces according to the structure proposed in the generic space—the mountain (along its peak and foot) is projected as a single element; the different days are projected into a single day; the moving individuals, however, are not fused, preserving their direction and relative positioning.6 Mental Spaces, Frames, Vital Relations

A conceptual integration network includes a generic space, input spaces, and a blended space; these are partial and dynamic ‘conceptual packets’ we construct as we think. The generic space contains general knowledge and concepts widely applicable to many thought processes. It is organized and structured by cognitive models (called ‘frames’) and is therefore abstract and not tied to any specific context or situation. Frames help us organize

214

Appendix VII: Conceptual Integration

and make sense of information by providing a basic set of expectations, assumptions, and associations—for instance, frames may draw on embodied knowledge such as image schemas, on interpersonal dynamics such as ‘competing’ or ‘flirting’, or on cultural constructs such as ‘justice’ or ‘love’. Input spaces emerge when activating the frames within the generic space. Unlike the generic space, input spaces can draw on knowledge from multiple domains to retrieve specific information and details, including anything from words and concepts to images and sensory experiences. The information across input spaces is linked via “vital relations”, the connective tissue that holds input spaces together. Vital relations are numerous, including ‘analogy’, ‘disanalogy’, ‘cause-effect’, ‘part-whole’, and ‘change’. Like other components of conceptual blends, vital relations are arbitrarily chosen and can dynamically change as our thoughts unfold. When we “run the blend”, attending to the generic space and its frames, and projecting elements within the input spaces according to specific vital relations, a blended space emerges. Nevertheless, the blended space is not simply a combination of the generic and input spaces, but a new structure that stems from their interaction—an emergent structure with unique properties not present in the source spaces, including new concepts, relationships, or thought patterns. Running the blend is an a-chronological process that entails: composition (i.e., constructing the image, like placing the two monks on a mountain); completion (i.e., supplying information or assumptions, like taking for granted that the monks are walking rather than crawling); elaboration (i.e., fleshing out the frames as they apply to the input spaces, like ‘monastic life’ and its connotations of meditation, prayer, and rituals); and integration (i.e., combining elements into a unified event, like the two monks walking on the same mountain at the same time). Neural Correlates of Conceptual Blending

Fauconnier and Turner (2002) suggest that conceptual blending is grounded in neural activity and claim that “in the neural interpretation of these cognitive processes, mental spaces are sets of activated neuronal assemblies, and the lines between elements correspond to coactivation-bindings of a certain kind” (p. 40). Although their claim is only speculative, we can interpret neuroscientific research to support it. For instance, at the neuronal level, conceptual blending may manifest itself as alterations in the patterns of synaptic connections (Calvin, 1996; Deacon, 2006; Phillips & Singer, 1997), as activations and subsequent integration of separate areas in the brain (Deacon, 2006), as synchronization of neuronal activation in terms of the level of activation and timing (Fotheringhame & Young, 1997; Phillips & Singer, 1997), as adaptations of neural circuits involved in motor activity, sensory modalities, and semantic memory (Deacon, 2006), and as new connections or activations between working memory and new information (Gernsbacher et al., 2001; Kintsch et al., 1999). Definitive neuroscientific research that supports conceptual blending is relatively scarce compared to the wealth of research that supports Conceptual Metaphor and Image Schema theories.7 Nonetheless, an emerging line of research supports conceptual blending within the ‘theoretical’ neurosciences, with studies that advance computational models that replicate a plausible mechanism at the neuronal level (e.g., Hodhod & Magerko, 2014; Thagard & Stewart, 2011).8

Appendix VII: Conceptual Integration

215

Conceptual Blending and Music

Music theorists and musicologists have drawn on the Conceptual Integration theory to formulate a broad framework for musical thinking and musical meaning; these studies cover a wide range of spheres, such as fleshing out correlations between music and text (Hsu & Su, 2014; Zbikowski, 1999, 2002), distilling the influence of programmatic elements (Tsougras & Stefanou, 2018), observing correlations between electroacoustic music and the emergence of meaning (Kendall, 2010), music and emotion (Spitzer, 2018), music and creativity (Brandt, 2021), and mapping the various cognitive spaces that partake in drawing meaning from musical multimedia (Atkinson, 2011; Chattah, 2006; Cook, 2017; Sayrs, 2003). Much of this work draws attention to structural parameters in the music and hence proposes ‘metaphorical’ correlations rather than ‘hypothetical scenarios’; in such cases, recurrent experiences of embodiment may motivate the abstract structure that links and relates mental spaces.9 Compared to Conceptual Metaphor theory, Conceptual Integration provides many advantages as an exploratory framework for film music, mainly where the music’s associations or connotations play a significant role.10 First, given that mental spaces are transient, as opposed to the more stable conceptual domains, conceptual blending allows us to disentangle the thought process that results in novel hypothetical scenarios based on associations or connotations stemming from the music or the narrative. Second, and related to the first point, associations or connotations entering the blend may or may not be conceived within the broader umbrella of embodied cognition or image schemas. Third, the resulting diagram provides an elegant framework to model the thought process unfolding during the blend. Fourth, while conceptual metaphors are based on projections between only two conceptual domains, conceptual blends allow for mappings between more than two mental spaces; in the case of films, it is critical to contemplate the projections among various mental spaces, including the music, the visuals, the narrative, the sound effects, and the dialogue, among others. Fifth, conceptual blending allows for a “recursive” process, one in which we may use a primary blend as an input space within a more encompassing blend; such recursive blending offers a valuable mechanism to explore multiple interpretative layers triggered by the music in a film.11 Notes 1. Conceptual Integration theory emerged as a reaction to well-established paradigms that “assume that natural language semantics can be adequately studied with the tools of formal logic” (Lakoff & Sweetser, 1994, p. ix). Early formulations of the theory date back to Fauconnier (1985); subsequent expansions on these early ideas include works by Dinsmore (1991) and Cutrer (1994). 2. Additionally, conceptual blending is akin to “multidimensional scaling”, which entails the coactivation of semantic spaces while positioning their components within a Euclidean space (Bauer, 2021; De Silva & Tenenbaum, 2004. 3. In comparing Conceptual Integration and Conceptual Metaphor theories, Zbikowski (2009) addresses the notion of mental spaces and conceptual domains, and notes that “the former [is] understood as ephemeral and pragmatic and the latter as relatively stable and abstract” (p. 370). This is one of the most immediate differences between the two theories. In addition, while metaphors rely on projections from one conceptual domain to another, conceptual blending relies on the fusion of two domains according to common abstract structures or patterns. Furthermore, conceptual blending is not concerned with grouping specific examples into larger, more inclusive categories; rather, it identifies unique instances of emergent patterns of similarity between cognitive domains. One fundamental similarity between these two theories, however, rests in their regarding metaphor as a conceptual rather than a purely linguistic phenomenon, one that relies

216

4. 5.

6. 7.

8.

9. 10.

11.

Appendix VII: Conceptual Integration

on projections among conceptual spaces. Antovic (2018) combines Conceptual Integration and Conceptual Metaphor theories to flesh out a theory of musical semantics. From Koestler’s (1964) The Act of Creation. Cited in Fauconnier and Turner (2002, p. 39). However, explaining or modeling the thought processes underlying the formation of metaphorical or hypothetical scenarios does not address the primary motivation for resorting to metaphors or hypothetical scenarios. Some research draws our attention to a potential motivation: the emotional underpinnings of the “AHA!” phenomenon that results from finding connections, solving a riddle, or ‘getting’ a metaphor (Bowden & Jung-Beeman, 2003; Thagard & Stewart, 2011). Thagard and Aubie (2008), for instance, propose that the “AHA!” moment activates the autonomous nervous system (i.e., heartbeat, sweating, etc.) and in turn an “emotional experience [arises] from a complex neural process that integrates cognitive appraisal of a situation with perception of internal physiological states” (p. 9). The blended space may involve elements that are not present in either input space (most often this is a replacement for a metonymic counterpart from one of the spaces). The primary challenge in supporting conceptual blending from a neural vantage point entails verifying potential mechanisms of “neuronal binding”, which would have further implications on symbolic representations and consciousness (Gibbs, 2001). In light of this challenge, Ritchie (2004) contends that “[it is] yet unknown how the information from these separate areas bound together . . . [and] there is no reason to think that such binding requires the replication of conceptual structure from each in a novel structure. Given the ability of most humans to spin out fanciful narratives, and to construct complex logical arguments, the creation of an entirely new ‘blended space’ for each conceptual integration would multiply conceptual representations in the brain to the point that memory capacity would quickly be exhausted” (p. 33). Nevertheless, within the broader challenge of explaining consciousness, researchers within theoretical neuroscience have proposed neurobiologically plausible models of neuronal binding based on synchronization of neural activity (Crick & Clark, 1994; Engel et al., 1999; Grandjean et al., 2008; van der Velde et al., 2017). Thagard and Stewart (2011) provide a computational-simulation account of neuronal binding as “the combination of previously unconnected mental representations constituted by patterns of neural activity” to support creative thinking, broadly understood to include “scientific discovery, technological invention, social innovation, and artistic imagination” (p. 3). To bridge the gap between theoretical neuroscience and embodiment, they propose a five-stage rationale that helps circumvent the problem of neural representation: “1) Creativity results from novel combinations of representations; 2) In humans, mental representations are patterns of neural activity; 3) Neural representations are multimodal, encompassing information that can be visual, auditory, tactile, olfactory, gustatory, kinesthetic, and emotional, as well as verbal; 4) Neural representations are combined by convolution, a kind of twisting together of existing representations; and 5) The causes of creative activity reside not just in psychological and neural mechanisms” (p. 2). Fauconnier and Turner (2002) account for such cases by including the “vital relations” of identity and analogy, which “apply across spaces to the topology of scales, image-schemas, and forcedynamic patterns inside mental spaces” (p. 105). Kövecses (2002) believes that conceptual metaphors are a special case within the much larger conceptual blending phenomenon proposed by Fauconnier and Turner. In fact, Fauconnier and Turner claim to have developed their theory to explain cognitive processes writ large, and not only those involving metaphorical statements—they refer to conceptual metaphors as “single-scope networks”. These advantages notwithstanding, Antović (2022) draws on both, the cognitive-linguistic theories of Conceptual Metaphor and Conceptual Integration, to formulate a “multilevel-grounded semantics”.

APPENDIX VIII Categorization

Categorization is fundamental to the human experience—we categorize the foods we eat, the pets we adopt, the movies we watch. Through categorization, we group similar objects or concepts together, which allows us to make predictions and generalizations about them. In this book, for instance, categorization has been vital in distinguishing the various sound design components (Chapter 2) or the musical topics weaved through the soundtrack (Chapter 9). In our cinematic experiences, just as in everyday life, categorization guides our actions, informs our perceptions, and determines our reactions to new experiences.1 The cognitive process of categorization is closely linked to representation, learning, and memory. In particular, the mechanisms whereby we construct or deduce categories are intricately connected to the way we interact with containers, and hence models of categorization implicitly draw on aspects of the CONTAINER schema: we conceive of categories in terms of their boundaries (rule-based model) or their contents (probabilistic model). In this appendix, I review these models, which have dominated the literature on human categorization, and set the space for our discussion of the cognitive processes involved in the perception of irony and related tropes, particularly as these involve film music.2 Categorization Approaches

Categories are, by and large, socially and culturally defined—the labels, category memberships, and associations brought to mind by categories are strongly influenced by our cultural environments and past experiences.3 Categories of film music topics, for instance, are defined by a large body of films and our experiences with those films. As this body of work accumulates, additional defining characteristics emerge and become emblematic of the category itself. Numerous models that explain the cognitive underpinnings of category construction have been developed within philosophy, psychology, the cognitive sciences, and related fields. Here, I review the two most influential approaches: rule-based and probabilistic.

218

Appendix VIII: Categorization

Rule-Based Approach

In his treatise Categories, Aristotle systematized an approach wherein category members feature singly necessary and jointly sufficient features—this approach is now referred to as “classical” or “rule-based”.4 Members of the category ‘pentagon’, as those shown in Figure VIII.1, for instance, must feature singly necessary and jointly sufficient conditions: (1) it must be a closed geometric form, (2) it must have five sides, and (3) its five internal angles must add up to 540 degrees. The absence of any one of these defining features disqualifies a shape from membership in the ‘pentagon’ category. A rule-based model works well for geometric figures and countless other categories, particularly because it allows us to formulate proto-logical expressions that simplify complex ideas, define the precise boundaries of categories, and apply uniformly to examples when testing for category membership.5 However, this model crumbles when approaching more fluid, context-based categorizations that reflect degrees of inclusion. Semantic (or artificial) categories present such challenges; concrete ones such as ‘furniture’ or ‘musical instrument’, and certainly more abstract ones such as ‘irony’, all defy attempts to define their necessary and sufficient features that delineate category boundaries. To address these shortcomings, Rosch and Mervis (1975) and Rosch (1978) built upon Wittgenstein’s (1953) notion of ‘family resemblance’ and further explored and systematized a probabilistic approach to categorization.

FIGURE VIII.1

Regular, irregular, and concave pentagons.

Probabilistic Approach

In his Philosophische Untersuchungen [Philosophical Investigations] (1953), Wittgenstein identifies artificial categories where members do not share necessary and sufficient features, but where members are instead related by overlapping attributes not common to all members—in the category ‘games’, for example, category members are related by overlapping (but not necessary or sufficient) attributes such as boards, cards, players, balls, and rules. Rosch and Mervis (1975) conducted a set of experiments that systematized a mode of categorization in which members illustrate variable and graded similarity relationships, setting up the foundation for the “exemplar” and the “prototype” models of categorization. In both the exemplar and prototype models, statistically prominent attributes determine category membership, and the degree of category membership of new examples depends on

Appendix VIII: Categorization

FIGURE VIII.2

219

Prototypicality of members of the ‘bird’ category.

the degree to which they exhibit those attributes; in both, also, new examples are compared to existing representations in memory. But whereas in the exemplar model we make category judgments by comparing new items to memory traces of individual category members (or ‘exemplars’), in the prototype model we make category judgments based on an abstract category member (or ‘prototype’) which exhibits the central tendencies of all category members.6 For instance, the exemplar model purports that in categorizing an animal as a ‘bird’, we compare it to existing memory traces of exemplars within the ‘bird’ category; in contrast, the prototype model purports that in categorizing an animal as a ‘bird’, we compare it to a single abstraction that subsumes the central characteristics of all members of the bird category.7 In later work, Rosch (1978) further develops the prototype model. She identifies that graded similarity to a prototype depends on the weights we give to statistically prominent attributes and to the values within attributes. Examples sharing a greater number of heavily weighted attributes and values are more typical members of the category. For the ‘bird’ category, for instance, the ‘mode of locomotion’ is one of the most heavily weighted attributes; and within the ‘mode of locomotion’ attribute, ‘flying’ is more heavily weighted than ‘running’ or ‘swimming’. Based on this weighting of attributes and values, wrens feature a higher degree of bird-like typicality than chickens, whereas penguins exhibit a lower degree of typicality than either chickens or wrens. Figure VIII.2 maps the possible cognitive mechanism

220

Appendix VIII: Categorization

that allows for the emergence of a prototype as a representative of a category featuring a high degree of typicality. Compared to the exemplar model, therefore, the prototype model offers the benefits of cognitive economy and ecological advantage—that is, averaging and weighting of attributes accounted for in the model provides us with “maximum information with the least cognitive effort” and allows us “not to differentiate one stimulus from others when that differentiation is irrelevant to the purposes at hand” (Rosch, 1978, p. 4). Rosch’s development of the prototype model took place at a time when embodied and ecological cognition gained traction and relevance.8 In defining possible attributes, for instance, Rosch extends beyond surface-level perceptual qualities and addresses the function of a category member in relation to perceivers and their intentions. That is, the contextual goals and action patterns afforded by the items categorized contribute to defining category membership. For example, given the right contextual goals and action patterns, a musical instrument can enter a different category—a piano can be a piece of furniture in the hands of an interior designer, or a weapon in the paws of Wile E. Coyote. Alternatively, any item, given the right contextual goals and action patterns, may be reframed within the musical instrument category—in the animated film The Triplets of Belleville, a newspaper, a refrigerator, a vacuum cleaner, and a bicycle wheel break free from their most common and salient categories, allowing the characters to reframe these objects as musical instruments and create a sound that challenges the very category of music.9 Given that the prototype model focuses on the contents rather than the boundaries of categories, Rosch (1978) attends to how categories are related hierarchically in terms of inclusiveness; she refers to this as the ‘vertical’ dimension. For instance, as shown in Figure VIII.3, the category ‘piano’ has subordinate categories such as ‘grand piano’ and the more specific, historic ‘fortepiano’. In turn, the category ‘piano’ belongs to superordinate categories such as ‘musical instrument’ and the even more comprehensive ‘instrument’ category. Situated toward the center level of inclusiveness, the category ‘piano’ is therefore the basic level.10 Rosch (1978) found that, when naming objects, individuals gravitate toward “the basic level, although they knew [the items’] correct superordinate and subordinate names” (p. 10) and that individuals are faster and more precise when drawing onto basic-level categories. In

FIGURE VIII.3

Vertical levels of inclusiveness.

Appendix VIII: Categorization

FIGURE VIII.4

221

Horizontal levels of inclusiveness.

the prior example, for instance, individuals would thus name the object ‘piano’, even when shown a picture of a grand piano. She argues that this tendency reflects a balance between efficiency (by minimizing the number of categories considered) and informativeness (by maximizing the most relevant information): calling an object “instrument” is highly efficient yet poorly informative; calling the same object a “fortepiano” is highly informative but not efficient; calling it a “piano” strikes the optimal balance between efficiency and effectiveness.11 In addition to a ‘vertical’ organizational principle based on levels of inclusiveness, Rosch contemplates a ‘horizontal’ dimension, which accounts for the segmentation of categories at a particular level, such as including ‘piano’, ‘timpani’, and ‘theremin’ within the ‘musical instrument’ superordinate category. (See Figure VIII.4.) She notes that in defining a category, individuals draw on prototypes that feature the most representative attributes of a category. In our example, ‘piano’ is more representative of the musical instrument category than timpani or theremin. Notes 1. An efficient categorization process provides an ecological and evolutionary advantage, as minor errors in classification may have drastic results; for instance, we categorize animals as snakes or hamsters so to infer attributes such as dangerous or harmless. 2. The neurological underpinnings of categorization are, to a great extent, unknown. Neuroscientific evidence suggests that we employ different cognitive mechanisms depending on our needs, circumstances, and capabilities. Farah and McClelland (1992), for instance, argue that categories are represented differently in the brain, whether these are the result of items’ surface-level features or the result of their function, particularly when items categorized could be understood as either ‘living’ or ‘nonliving’—that is, categorization of living items is strongly dependent on surface-level features, whereas categorization of nonliving items is strongly dependent on their function. Smith et al. (2016) provide a different perspective, noting “the presence in human and primate brains of multiple, dissociable processes within the overall category-learning system” (p. 13) and asserting that we engage different cognitive strategies for categorization depending on the ‘density’ of categories, resorting to the exemplar model for sparse categories and to the prototype model for densely populated categories. 3. Lakoff’s (1987) principle of ‘domain-of-experience’ extends the notion of culturally defined categories even further, to identify categories in which all its members are related by a cultural or collective experience; his most telling example is the category ‘balan’ in the Dyirbal language, which includes “women”, “fire”, and “dangerous things”. 4. Aristotle argued that every object or concept belongs to a specific category or genus, and that this category can be determined by examining the defining characteristics or essential properties

222

5. 6.

7.

8.

9. 10.

11.

Appendix VIII: Categorization

of the object or concept. This approach, also known as “essentialism”, became a cornerstone of Aristotle’s and his followers’ philosophies. Because of these benefits, this approach is optimal from an evolutionary perspective, as it allows us to make rapid inferences or predictions based on partial information or a subset of features (Medin & Coley, 1998). Rosch (1978) characterizes prototypes as “those members of a category that most reflect the redundancy structure of the category as a whole” (p. 12). However, because a prototype does not correspond to an actual instance representative of a category, a prototype could be qualified as an average or central tendency of all category members (Kruschke, 2008). Recently, the exemplar model has fallen out of favor because of its weaknesses related to memory and information retrieval, which may result in ecological and adaptive risks. The prototype model, instead, offers the ecological advantage of allowing for category judgments based on a single abstract prototype, rather than on a number of exemplars stored in memory. That said, Rosseel (2002) considers the possibility of multiple prototypes per category. These tendencies toward ecological and evolutionary cognition were pervasive within the scholarly community of the time, yet scholars seldom acknowledged each other’s work. Rosch (1978) notes that “given an actor with the motor programs for sitting, it is a fact of the perceived world that objects with the perceptual attributes of chairs are more likely to have functional sit-on-able-ness than objects with the appearance of cats” (p. 4). Additionally, from an ecological perspective, she notes that the perception of certain attributes is species-specific; for instance, because a “dog’s sense of smell is more highly differentiated than a human’s, [the] structure of the world for a dog must surely include attributes of smell that we, as a species, are incapable of perceiving . . . Furthermore, because a dog’s body is constructed differently from a human’s, its motor interactions with objects are necessarily differently structured” (p. 4). Although not directly cited within her work, these passages are a distinct nod to Gibson’s notion of affordances. In turn, in further developing his notion of affordances, Gibson (1979) draws on Rosch’s notion of prototypes, albeit without explicitly citing her. He notes, for instance, that “the fact that a stone is a missile does not imply that it cannot be other things as well. It can be a paperweight, a bookend, a hammer, or a pendulum bob. It can be piled on another rock to make a cairn or a stone wall. These affordances are all consistent with one another. The differences between them are not clear-cut, and the arbitrary names by which they are called [i.e., categories] do not count for perception” (p. 134). Gibson then adds that “The theory of affordances rescues us from the philosophical muddle of assuming fixed classes of objects, each defined by its common features and then given a name . . . You do not have to classify and label things in order to perceive what they afford” (p. 134). See The Triplets of Belleville. [00:55:10] In addition, category members at the basic level may activate motor programs and modal representations more vividly than at the more abstract and disembodied superordinate levels (Grodal, 2009). For instance, the basic level ‘piano’ elicits a more immediate response during retrieval than the more abstract superordinate level ‘instrument’. The balance between efficiency and effectiveness notwithstanding, situational or contextual variables prompt different levels of categorization. For instance, an orchestra conductor addressing the woodwind performers would opt for the subordinate level ‘piccolo flute’ to the basic level ‘flute’; in such cases, although addressing the subordinate level demands more detailed processing, it helps avoid grave misunderstandings.

APPENDIX IX Additional Examples

In this last appendix, I offer additional examples that may be of interest to readers. To contextualize the music within the film’s action and storyline, a summary of the film and the scene introduces each example.1 Following those brief introductions, I offer questions and discussion points to draw the reader’s attention to salient musical facets. However, there is no single or correct answer to any of the questions and discussion points; instead, these will prompt further reflection and give readers an opportunity to apply the ESMAMAPA framework in exploring their own sensitivity and intuitions about film music’s power to shape our experiences and interpretations. Last, while the cognitive and semiotic mechanisms the ESMAMAPA framework encompasses may prove sufficient when unpacking most interpretations, I include questions at the end of the appendix that will prompt the reader to test this framework’s explanatory limits and underlying premises. 1917

Several devastating years into World War I, two British lance corporals—Schofield and Blake—race against time, crossing over into enemy territory to deliver a message that could save thousands, including Blake’s brother. In a scene early in the film, the two soldiers embark on their journey through no-man’s-land, first reaching abandoned German trenches. • How does the music reflect the psychological tension leading to the discovery that the trenches have been abandoned? • As the soldiers cross through the mud, they glimpse corpses trapped in barbed wire. How does the music allow us access to the characters’ psychological apprehension? • What effect does the final cadential gesture in this music cue have on our interpretation of the storyline? 45 Years

Kate and Geoff plan to celebrate their 45th wedding anniversary with many friends. Weeks before the celebration, Geoff receives a letter notifying him that his first love, Katya, was

224

Appendix IX: Additional Examples

discovered after many years, frozen in the glaciers in Switzerland. As the party nears, Kate finds pictures that Geoff kept from his time with Katya in the Swiss Alps and learns that Katya was pregnant when she died. In the film’s last scene, during the celebration, Kate and Geoff dance to their wedding song: “Smoke Gets in Your Eyes” by The Platters. • How do the lyrics of the song color our perception of their marriage and their future as a couple? • What other musical elements contribute to this perception? • How do other music cues in the film help delineate Kate’s emotional journey? A.I.

It is the twentieth-second century. Natural disasters stemming from global climate change have reduced the world’s population. Henry and Monica, whose child is in suspended animation because of a rare disease, adopt David, a robotic boy capable of experiencing love. In an early scene, Monica struggles to accept David as her child and tests his (and her own) boundaries of adaptive human behavior. • Why does the music introduce musical gestures predominantly in the high register? • How does the tension between David’s human appearance and his/its mechanical construction play within the music’s timbres? • The music introduces dissonant undertones within an otherwise innocent, childlike texture. What is the purpose of these dissonances? • How does the cadential gesture at the end of the scene influence our perception of the storyline? Avengers: Endgame

After Thanos erased half of all life and decimated the universe, the Avengers plan to use the Infinity Stones to reverse his actions. In the final scene, Captain America, Thor, and Iron Man cannot take control of Thanos’s brutal forces. In defeat, Captain America seems to surrender. At that moment, a portal opens. Okoye, T’Challa, and Shuri appear, and then Falcon flies out. Other portals open, bringing Dr. Strange, Spiderman, other Avengers, the Guardians of the Galaxy, the Ravagers, and the armies of Wakanda and Asgard. At Captain America’s command, “Avengers: assemble!”, they join forces to fight Thanos and his army. • As Okoye, T’Challa, and Shuri appear, what instrumental choices bring about thoughts of honor and courage? • How does the music reflect the ever-strengthening coalescing forces as superheroes arrive? • Although the music initially projects an ametric structure, it gradually defines a duple meter. How does the music’s temporality influence our perception of the events to unfold? Beast

In a remote community coping with a string of unsolved rapes and murders of young girls, troubled 27-year-old Moll falls for a mysterious outsider. As their relationship blossoms, he

Appendix IX: Additional Examples

225

empowers her to escape her wealthy yet oppressive family. He comes under suspicion for the rapes and murders, but she defends him. In the film’s opening scene, Moll sings in a choir conducted by her mother; interspersed, we see flashes of some of the missing girls’ obituaries. In a later scene, the police detain Moll and bring her to their headquarters for interrogation. • In the first scene, the choir piece’s idyllic, perfectly in-tune setting becomes increasingly distorted with emergent low, noisy, dissonant rumblings. What is the music conveying at this moment in the film? How does the music prompt us to construct that interpretation? • In the latter scene, the music blends with the sound of Moll’s heartbeat. Although her heart’s pace remains steady, the loudness noticeably increases. How does the soundtrack influence us viscerally at this moment in the film? Cassandra’s Dream

The Blaine brothers, Ian and Terry, enjoy a life of good fortune. Their luck soon changes when they purchase a luxury sailboat at an oddly low price and name it “Cassandra’s Dream”, unaware of the mythological antecedents of the name. • What effect does the minimalist-style score have on our perception of the characters and the narrative? • How does the inconclusive nature of the cadential gestures to (nearly) all cues resonate with the film’s central message? • What is the combined effect of the minimalist-style score and the inconclusive gestures on either realizing or thwarting our musical (and dramatic) expectations? Chicken Run

The lives of (anthropomorphic) egg-laying chickens are doomed when a chicken-pie-making machine arrives at the chicken farm—they must organize and escape. The main title sequence illustrates a series of unsuccessful attempts at breaking free—unable to fly, the chickens try some bizarre strategies. • What topic do the surface-level musical features suggest? What is the function of such a topic during the main title sequence? • Toward the end of the main title sequence, there is a sudden shift to a triple meter. How does this shift modulate our perception of that moment in the film? • What type of cadence concludes the main title sequence? To what effect? Deadpool

Wade Wilson, a former Special Forces operative, becomes romantically involved with Vanessa. Diagnosed with terminal cancer, Wade vanishes to save Vanessa from heartache. He submits himself to Ajax, a mad scientist, who misleadingly promises to heal him and injects him with a serum to awaken his mutant genes. The experiment disfigures and transforms Wade into Deadpool, yet leaves him with enhanced healing powers, which

226

Appendix IX: Additional Examples

he uses to hunt down Ajax. In the film’s last scene, Deadpool saves Vanessa from a collapsing ship. The self-healing but disfigured superhero apologizes to his former girlfriend for abandoning her, and although she still has trouble with Wade, she is willing to move forward with Deadpool. • What is the role of the ‘sexophone’ topic in the scene? Does it function primarily to support the narrative, indicate the film’s genre, or define the character’s identity? • As Vanessa forgives Wade and acknowledges that she will eventually get used to his hideous face, how does the sound design reflect his taking control of the narrative? Midsommar

Dani and Christian travel to Sweden to visit their friend’s rural hometown during its midsummer festivities. Their idyllic retreat is soon transformed by a series of violent and uncanny rituals led by a Scandinavian pagan cult. In an early scene, Dani awakens from a strange psychedelic mushroom trip and joins others in their hike toward the village. Upon arrival, the village’s members greet and welcome them to the celebrations with music, food, and gifts. • A repeating, harmonically static upward gesture on flutes underscores the long journey. How does the repetition of this gesture impact us? • How does the music’s metrical structure influence our perception of time? • As the group arrives, a flute ensemble welcomes them, prompting a transference in the diegesis. How does this transference subliminally prepare us for the events to unfold? Mission Impossible II

Ethan Hunt leads the Impossible Mission Force team on a quest to seize a deadly virus before it reaches the hands of terrorists. In an early scene, Ethan travels to Seville, Spain, to recruit Nyah Nordoff-Hall, a skillful thief and the chief terrorist’s former girlfriend; he spots her at a private gathering, stealing a necklace. In the film’s last scene, Nyah is cleared of her criminal record and joins Ethan for a vacation in Sydney. • What topic does the Flamenco guitar suggest in the first scene during the private gathering? • The guitar melody will become their leitmotif. How do the meter, modality, and other musical parameters reflect the characters’ relationship? • How does the music map the slow-motion in the visuals as they first cross paths? • What are the most significant leitmotif transformations in the ending scene? What meaning(s) do these transformations convey, and how do these convey such meaning(s)? Saving Private Ryan

In this film, Captain John Miller leads his squad behind enemy lines to rescue Private James Ryan. In the film’s first scene, an aging veteran walks through a cemetery and becomes overwhelmed with emotions as he recalls his time as a soldier. The film’s last scene reveals that

Appendix IX: Additional Examples

227

the elderly veteran is Ryan, filled with feelings of grief and gratitude while saluting Miller’s gravestone. • How do the music’s cadential gestures in the first and last scenes help us embark on the characters’ journeys and subsequently bring that journey to a conclusion? • What instrumental choices bring about thoughts of honor, decency, and courage? The Hunger Games: Mockingjay – Part I

The story in the third installment of the film series continues to follow Katniss Everdeen, who survived the Hunger Games twice. She becomes a symbol for mass uprising and coordinates to rescue Peeta, one of her love interests, from the oppressive forces of the Capitol. In the scene, Katniss sings a rallying cry she learned from her father. Each verse of this strophic song repeats an initial question—“Are you, are you, coming to the tree?”—then intersperses four events—“They strung up a man, they say who murdered three”, “Where dead man called out, for his love to flee”, “Where I told you to run, so we’d both be free”, “Wear a necklace of hope, side by side with me”—and closes all verses by repeating an implicit proposition—“Strange things did happen here, no stranger would it be, if we met at midnight, in the hanging tree.” As the music continues, masses of protesters join in singing the anthem while launching an attack to destroy a hydroelectric dam, the Capitol’s primary source of energy. • Although the song initially feels like an innocent nursery rhyme, its lyrics project a rebel anthem. How does this tension play out within the music and the film’s context? • How do the song’s Appalachian flavor, duple meter, and increasing loudness and number of voices impact our understanding of the scene? • Katniss’s diegetic singing smoothly unfolds into a non-diegetic orchestral rendition of the song. What sound design technique is at play? What are the narrative and interpretational entailments of using this technique at this moment in the film? The Lobster

In a dystopian world, single people have a month and two weeks to find a partner, or they are converted into a beast of their choice and freed into the wild. David, who has recently become single after a 12-year relationship, is eager to find a new partner, but realizing he has no outstanding physical traits, he proactively chooses to become a lobster. In a scene, the Greek song “Ti einai afto’ pou to lene agape” underscores slow-motion visuals of a ritual in which single individuals hunt each other in the woods. The lyrics to the first verse translate to “What is it that is called love? What is that? What drives the heart secretly, and who feels that nostalgia?” • How do the song and its lyrics work in tandem with (or against) the visuals to enhance the tension between their and our conception of love? Up

Carl and Ellie fantasize about escaping life’s hurdles by restoring an abandoned house and moving it to a mountain peak overlooking Paradise Falls. The opening sequence, which

228

Appendix IX: Additional Examples

contains only music and no dialogue or sound effects, spans their entire life together, from marriage to rebuilding the house, her suffering a miscarriage, saving for a trip to Paradise Falls but repeatedly spending those savings on more pressing needs, and her falling ill and untimely death. • How do the piece’s ‘music box’ sound and other timbres influence our interpretation? • Balloons are a vital element in the story. How is their movement (up and down) reflected in the score? • The montage spans a great number of moments in their married life. How does the music help delimit these moments? • Although much of the montage shows them working, how does the music’s meter infuse a more playful tone? • How are the events, such as Ellie’s suffering after her miscarriage or Carl’s mourning of Ellie’s death, reflected in the music? Questions for Further Study

• Empathy: Have you encountered film music moments that elicit an empathic response through a mechanism other than subvocalization, entrainment, or contagion? • Schemas and Metaphor: How would the CYCLE or the PART-WHOLE schema manifest itself through the music? What impact would mapping these schemas onto the musical fabric have on our interpretation of a scene or the narrative? • Affordances: Can you find an example of a musical parameter (other than meter and tonal centricity) that delimits our interaction with the environment and thereby defines our interpretation of a scene? • Memory and Auditory Perception: Can you find an instance of a film that establishes a leitmotif by means other than drawing on hardwired responses, musical affordances, acoustic resemblances, or musical archetypes? • Archetypes: Have you come across a new film music topic? Can you trace (or speculate about) the source of that topic and identify its surface-level markers? • Context: Are there film moments in which the music supplies a contradictory subtext that complicates the narrative? • Holistic: Are there interpretations that resist explanation through the lens of the ESMAMAPA framework? Note 1. Note to instructors: The Copyright Act allows for showing copyrighted films or clips during regular classroom instruction, and only to students in the course (thus excluding ‘public’ viewings). Instructors must use a legal copy of the film and may not charge a fee to the students.

REFERENCES

Abril, C. (2001). The use of labels to describe pitch changes by bilingual children. Bulletin of the Council for Research in Music Education, 151, 31–40. Acitores, A. P. (2011). Towards a theory of proprioception as a bodily basis for consciousness in music. Music and Consciousness: Philosophical, Psychological, and Cultural Perspectives, 1, 215. Adolph, K. E., Eppler, M. A., & Gibson, E. J. (1993). Crawling versus walking infants’ perception of affordances for locomotion over sloping surfaces. Child Development, 64(4), 1158–1174. Agawu, V. K. (1991). Playing with signs: A semiotic interpretation of classic music. Princeton University Press. Agosta, L. (2014). A rumor of empathy: Rewriting empathy in the context of philosophy. Springer. Agus, T. R., Thorpe, S. J., Suied, C., & Pressnitzer, D. (2010, May). Characteristics of human voice processing. In Proceedings of 2010 IEEE international symposium on circuits and systems (pp. 509– 512). IEEE. Alger, S. E., Kensinger, E. A., & Payne, J. D. (2018). Preferential consolidation of emotionally salient information during a nap is preserved in middle age. Neurobiology of Aging, 68, 34–47. Anderson, M. L. (2010). Neural reuse: A fundamental organizational principle of the brain. Behavioral and Brain Sciences, 33(4), 245–266. Anderson, M. L. (2016). Précis of after phrenology: Neural reuse and the interactive brain. Behavioral and Brain Sciences, 39. Anishchenko, V. S., Balanov, A. G., Janson, N. B., Igosheva, N. B., & Bordyugov, G. V. (2000). Entrainment between heart rate and weak noninvasive forcing. International Journal of Bifurcation and Chaos, 10(10), 2339–2348. Antović, M. (2018). From expectation to concepts: Toward multilevel grounding in musical semantics. Cognitive Semiotics, 9(2), 105–138. Antović, M. (2022). Form, meaning and intentionality: The case of metaphor in music. In Metaphors and analogies in sciences and humanities: Words and worlds (pp. 553–577). Springer International Publishing. Asch, S. E., & Nerlove, H. (1960). The development of double function terms in children. In H. Wapner & B. Kaplan (Eds.), Perspectives in psychological theory (pp. 41–60). International Universities Press. Atkinson, R. C., & Shiffrin, R. M. (1968). Human memory: A proposed system and its control processes. In Psychology of learning and motivation (Vol. 2, pp. 89–195). Academic Press. Atkinson, S. (2011). Canons, augmentations, and their meaning in two works by Steve Reich. Music Theory Online, 17(1).

230

References

Aurnague, M., Hickmann, M., Vieu, L., & Shtalbi, H. (2007). The categorization of spatial entities in language and cognition (Human Cognitive Processing). John Benjamins. Aziz-Zadeh, L., Iacoboni, M., Zaidel, E., Wilson, S., & Mazziotta, J. (2004). Left hemisphere motor facilitation in response to manual action sounds. European Journal of Neuroscience, 19(9), 2609–2612. Bach, P., Nicholson, T., & Hudson, M. (2014). The affordance-matching hypothesis: How objects guide action understanding and prediction. Frontiers in Human Neuroscience, 8, 254. Baddeley, A. D., & Hitch, G. (1974). Working memory. In Psychology of learning and motivation (Vol. 8, pp. 47–89). Academic Press. Baker, D. J., & Müllensiefen, D. (2017). Perception of leitmotives in Richard Wagner’s Der Ring des Nibelungen. Frontiers in Psychology, 8, 662. Balkwill, L. L., & Thompson, W. F. (1999). A cross-cultural investigation of the perception of emotion in music: Psychophysical and cultural cues. Music Perception, 17(1), 43–64. Bargh, J. A., & Chartrand, T. L. (1999). The unbearable automaticity of being. American Psychologist, 54(7), 462. Bargh, J. A., Williams, L., Huang, J., Song, H., & Ackerman, J. (2010). From the physical to the psychological: Mundane experiences influence social judgment and interpersonal behavior. Behavioral and Brain Sciences, 33(4), 267–268. Barsalou, L. W. (2008). Cognitive and neural contributions to understanding the conceptual system. Current Directions in Psychological Science, 17(2), 91–95. Barsalou, L. W., Simmons, W. K., Barbey, A. K., & Wilson, C. D. (2003). Grounding conceptual knowledge in modality-specific systems. Trends in Cognitive Sciences, 7(2), 84–91. Barthes, R. (1985). The responsibility of forms: Critical essays on music, art, and representation. University of California. Bartlett, F. C. (1933). Remembering: A study in experimental and social psychology. British Journal of Educational Psychology, 3(2), 187–192. Bauer, A. (2021). Tone-color, movement, changing harmonic planes: Cognition, constraints and conceptual blends in modernist music (eScholarship). University of California. Bazelon, I. (1975). Knowing the score: Notes on film music. Van Nostrand Reinhold Company. Bell, D. A. (1994). Getting the best score for your film: A filmmakers’ guide to music scoring. SilmanJames Press. Belletto, S. (2008). “Cabaret” and antifascist aesthetics. Criticism, 50(4), 609–630. Berman, M. G., Jonides, J., & Lewis, R. L. (2009). In search of decay in verbal short-term memory. Journal of Experimental Psychology: Learning, Memory, and Cognition, 35(2), 317. Bestelmeyer, P. E., Maurage, P., Rouger, J., Latinus, M., & Belin, P. (2014). Adaptation to vocal expressions reveals multistep perception of auditory emotion. Journal of Neuroscience, 34(24), 8098–8105. Bey, C., & McAdams, S. (2002). Schema-based processing in auditory scene analysis. Perception & Psychophysics, 64(5), 844–854. Biancorosso, G. (2010). The shark in the music. Music Analysis, 29(1–3), 306–333. Biancorosso, G. (2013). Memory and the leitmotif in cinema. In Representation in western music. Cambridge University Press. Blood, A. J., Zatorre, R. J., Bermudez, P., & Evans, A. C. (1999). Emotional responses to pleasant and unpleasant music correlate with activity in paralimbic brain regions. Nature Neuroscience, 2(4), 382–387. Boltz, M., Schulkind, M., & Kantra, S. (1991). Effects of background music on the remembering of filmed events. Memory & Cognition, 19(6), 593–606. Boltz, M. G. (2001). Musical soundtracks as a schematic influence on the cognitive processing of filmed events. Music Perception, 18(4), 427–454. Boltz, M. G. (2004). The cognitive processing of film and musical soundtracks. Memory & Cognition, 32(7), 1194–1205.

References

231

Boltz, M. G., Ebendorf, B., & Field, B. (2009). Audiovisual interactions: The impact of visual information on music perception and memory. Music Perception, 27(1), 43–59. Bourne, J. (2021). Hearing film music topics outside the movie theatre: Listening cinematically to pastorals. In The Oxford handbook of cinematic listening. Oxford University Press. Bowden, E. M., & Jung-Beeman, M. (2003). Aha! Insight experience correlates with solution activation in the right hemisphere. Psychonomic Bulletin & Review, 10(3), 730–737. Brandt, A. (2021). Defining creativity: A view from the arts. Creativity Research Journal, 33(2), 81–95. Brandt, L. (2009). Metaphor and the communicative mind. Cognitive Semiotics, 5(1–2), 37–107. Bregman, A. (1990). Auditory scene analysis: The perceptual organization of sound. MIT Press. Bremner, A. J., Caparos, S., Davidoff, J., de Fockert, J., Linnell, K. J., & Spence, C. (2013). ‘Bouba’ and ‘Kiki’ in Namibia? A remote culture makes similar shape–sound matches, but different shape–taste matches to westerners. Cognition, 126, 165–172. Bribitzer-Stull, M. (2015). Understanding the leitmotif. Cambridge University Press. Brochard, R., Abecasis, D., Potter, D., Ragot, R., & Drake, C. (2003). The “ticktock” of our internal clock: Direct brain evidence of subjective accents in isochronous sequences. Psychological Science, 14(4), 362–366. Brodsky, W., Kessler, Y., Rubinstein, B. S., Ginsborg, J., & Henik, A. (2008). The mental representation of music notation: Notational audiation. Journal of Experimental Psychology: Human Perception and Performance, 34(2), 427. Browne, R. (1981). Tonal implications of the diatonic set. In Theory Only, 5(6), 3–21. Buccino, G., Riggio, L., Melli, G., Binkofski, F., Gallese, V., & Rizzolatti, G. (2005). Listening to actionrelated sentences modulates the activity of the motor system: A combined TMS and behavioral study. Cognitive Brain Research, 24(3), 355–363. Buccino, G., Vogt, S., Ritzl, A., Fink, G. R., Zilles, K., Freund, H. J., & Rizzolatti, G. (2004). Neural circuits underlying imitation learning of hand actions: An event-related fMRI study. Neuron, 42(2), 323–334. Buhler, J. (2016). Branding the Franchise: Music, Opening Credits, and the (Corporate) Myth of Origin. In Music in Epic Film (pp. 17–40). Routledge. Bullerjahn, C., & Güldenring, M. (1994). An empirical investigation of effects of film music using qualitative content analysis. Psychomusicology: A Journal of Research in Music Cognition, 13(1–2), 99. Bundgaard, P. F. (2009). Are cross-domain mappings psychologically deep, but conceptually shallow? What is still left to test for conceptual metaphor theory. Cognitive Semiotics, 5(1–2), 400–407. Burianova, H., McIntosh, A. R., & Grady, C. L. (2010). A common functional brain network for autobiographical, episodic, and semantic memory retrieval. Neuroimage, 49(1), 865–874. Butler, D. (1989). Describing the perception of tonality in music: A critique of the tonal hierarchy theory and a proposal for a theory of intervallic rivalry. Music Perception, 6(3), 219–241. Byron, T. P. (2008). The processing of pitch and temporal information in relational memory for melodies [PhD dissertation, University of Western Sydney]. Callan, D. E., Tsytsarev, V., Hanakawa, T., Callan, A. M., Katsuhara, M., Fukuyama, H., & Turner, R. (2006). Song and speech: Brain regions involved with perception and covert production. Neuroimage, 31(3), 1327–1342. Calvin, W. H. (1996). The cerebral code: Thinking a thought in the mosaics of the mind. MIT Press. Cambouropoulos, E. (2001). Melodic cue abstraction, similarity, and category formation: A formal model. Music Perception, 18(3), 347–370. Cantor, J. (2004). “I’ll never have a clown in my house”—why movie horror lives on. Poetics Today, 25(2), 283–304. Carr, L., Iacoboni, M., Dubeau, M. C., Mazziotta, J. C., & Lenzi, G. L. (2003). Neural mechanisms of empathy in humans: A relay from neural systems for imitation to limbic areas. Proceedings of the National Academy of Sciences, 100(9), 5497–5502. Challis, B. H., Velichkovsky, B. M., & Craik, F. I. (1996). Levels-of-processing effects on a variety of memory tasks: New findings and theoretical implications. Consciousness and Cognition, 5(1–2), 142–164. Chao, L. L., & Martin, A. (2000). Representation of manipulable man-made objects in the dorsal stream. Neuroimage, 12(4), 478–484.

232

References

Chattah, J. (2006). Semiotics, pragmatics, and metaphor in film music analysis [Unpublished PhD dissertation, Florida State University]. Chattah, J. (2008). Conceptual integration and film music analysis. Semiotics, 772–783. Chattah, J. (2015). Film music as embodiment. In M. Coëgnarts & P. Kravanja (Eds.), Embodied cognition and cinema. Leuven University Press. Cherlin, M. (2017). Varieties of musical irony. Cambridge University Press. Choi, S., McDonough, L., Bowerman, M., & Mandler, J. (1999). Early sensitivity to language-specific spatial categories in English and Korean. Cognitive Development, 14(2), 241–268. Cisek, P. (2007). Cortical mechanisms of action selection: The affordance competition hypothesis. Philosophical Transactions of the Royal Society B: Biological Sciences, 362(1485), 1585–1599. Clark, A. (1989). Microcognition: Philosophy, cognitive science, and parallel distributed processing. MIT Press. Clarke, E. (2005). Ways of listening: An ecological approach to the perception of musical meaning. Oxford University Press. Clynes, M. (1977). Sentics: The touch of emotions. Anchor Press. Cobo, R. M. D. (2003). Parody and satire in Burgess’ a clockwork orange and in Kubrick’s cinematic adaptation. Estudios Humanísticos. Filología, (25), 57–69. Coëgnarts, M. (2019). Film as embodied art: Bodily meaning in the cinema of Stanley Kubrick. Academic Studies Press. Coëgnarts, M., & Kravanja, P. (2012). Towards an embodied poetics of cinema: The metaphoric construction of abstract meaning in film. Alphaville, 4, 1–18. Coëgnarts, M., & Kravanja, P. (2015). With the past in front of the character: Evidence for spatialtemporal metaphors in cinema. Metaphor and Symbol, 30(3), 218–239. Cohen, A. J. (2014). Film music from the perspective of cognitive science. In The Oxford handbook of film music studies (pp. 96–130). Oxford University Press. Cohen, M. E., & Carr, W. J. (1975). Facial recognition and the von Restorff effect. Bulletin of the Psychonomic Society, 6(4), 383–384. Cohn, R. (2012). Audacious euphony: Chromatic harmony and the triad’s second nature. Oxford University Press. Collier, W. G., & Hubbard, T. L. (2001). Musical scales and evaluations of happiness and awkwardness: Effects of pitch, direction, and scale mode. American Journal of Psychology, 114(3), 355–375. Cook, N. (1998). Analysing musical multimedia. Clarendon Press. Cook, N. (2017). Music, performance, meaning: Selected essays. Routledge. Cox, A. (1999). The metaphoric logic of musical motion and space [Unpublished PhD dissertation, University of Oregon]. Cox, A. (2001). The mimetic hypothesis and embodied musical meaning. Musicae Scientiae, 5(2), 195–212. Cox, A. (2016). Music and embodied cognition: Listening, moving, feeling, and thinking. Indiana University Press. Craik, F. I., & Lockhart, R. S. (1972). Levels of processing: A framework for memory research. Journal of Verbal Learning and Verbal Behavior, 11(6), 671–684. Crick, F., & Clark, J. (1994). The astonishing hypothesis. Journal of Consciousness Studies, 1(1), 10–16. Crowder, R. G., Reznick, J. S., & Rosenkrantz, S. L. (1991). Perception of the major/minor distinction: V. Preferences among infants. Bulletin of the Psychonomic Society, 29(3), 187–188. Cutrer, L. M. (1994). Time and tense in narrative and in everyday language [Doctoral dissertation, University of California, San Diego]. D’Aloia, A. (2012). The intangible ground – A neurophenomenology of the film experience. NECSUS: European Journal of Media Studies, 1(2), 219–239. Deacon, T. (2006). The aesthetic faculty. In The artful mind: Cognitive science and the riddle of human creativity (pp. 21–53). Oxford University Press. Declerck, G. (2013). Why motor simulation cannot explain affordance perception. Adaptive Behavior, 21(4), 286–298.

References

233

Deliège, I. (1992). Recognition of the Wagnerian leitmotiv: Experimental study based on an excerpt from Das Rheingold. Musik Psychologie, 9, 25–54. Deliège, I. (2001). Prototype effects in music listening: An empirical approach to the notion of imprint. Music Perception, 18(3), 371–407. Deliège, I., & Mélen, M. (1997). Cue abstraction in the representation of musical form. In Perception and cognition of music (pp. 374–397). Psychology Press. Deliège, I., Mélen, M., Stammers, D., & Cross, I. (1996). Musical schemata in real-time listening to a piece of music. Music Perception, 14(2), 117–159. Deroy, O., & Auvray, M. (2013). A new Molyneux’s problem: Sounds, shapes and arbitrary crossmodal correspondences. In O. Kutz, M. Bhatt, S. Borgo, & P. Santos (Eds.), Second international workshop the shape of things (pp. 61–70). International Association of Ontology and its Applications. De Silva, V., & Tenenbaum, J. B. (2004). Sparse multidimensional scaling using landmark points (Vol. 120, Technical report). Stanford University. D’Esposito, M. (2007). From cognitive to neural models of working memory. Philosophical Transactions of the Royal Society B: Biological Sciences, 362(1481), 761–772. De Souza, J. (2017). Music at hand: Instruments, bodies, and cognition. Oxford University Press. De Waal, F. B. (2007). The ‘Russian doll’ model of empathy and imitation. In On being moved: From mirror neurons to empathy (pp. 35–48). John Benjamins Publishing Company. Dewell, R. (2005). Dynamic patterns of CONTAINMENT. In B. Hampe & J. Grady (Eds.), From perception to meaning: Image schemas in cognitive linguistics (pp. 369–394). Walter de Gruyter GmbH. de Wit, M. M., de Vries, S., van der Kamp, J., & Withagen, R. (2017). Affordances and neuroscience: Steps towards a successful marriage. Neuroscience & Biobehavioral Reviews, 80, 622–629. Dimberg, U., & Öhman, A. (1996). Behold the wrath: Psychophysiological responses to facial stimuli. Motivation and Emotion, 20(2), 149–182. Dimberg, U., Thunberg, M., & Elmehed, K. (2000). Unconscious facial reactions to emotional facial expressions. Psychological Science, 11(1), 86–89. Dinsmore, J. (1991). Partitioned representations. In Partitioned representations (pp. 45–91). Springer. Di Stefano, N., Vuust, P., & Brattico, E. (2022). Consonance and dissonance perception. A critical review of the historical sources, multidisciplinary findings, and main hypotheses. Physics of Life Reviews, 43, 273–304. Doelling, K. B., & Poeppel, D. (2015). Cortical entrainment to music and its modulation by expertise. Proceedings of the National Academy of Sciences, 112(45), E6233–E6242. Doll, C. (2018). Was it diegetic, or just a dream? Music’s paradoxical place in the film INCEPTION. Society for Music Theory: Videocast Journal (SMT-V), 4(1). Donnelly, K. J. (1998). The classical film score forever? Batman, Batman Returns and post-classical film music. Contemporary Hollywood Cinema, 142–155. Dowling, W. J. (1972). Recognition of melodic transformations: Inversion, retrograde, and retrograde inversion. Perception & Psychophysics, 12(5), 417–421. Dowling, W. J. (1978). Scale and contour: Two components of a theory of memory for melodies. Psychological Review, 85(4), 341. Dowling, W. J., & Bartlett, J. C. (1981). The importance of interval information in long-term memory for melodies. Psychomusicology: A Journal of Research in Music Cognition, 1(1), 30. Dowling, W. J., & Fujitani, D. S. (1971). Contour, interval, and pitch recognition in memory for melodies. The Journal of the Acoustical Society of America, 49(2B), 524–531. Dyson, M. C., & Watkins, A. J. (1984). A figural approach to the role of melodic contour in melody recognition. Perception & Psychophysics, 35(5), 477–488. Eitan, Z. (2013). How pitch and loudness shape musical space and motion. In The psychology of music in multimedia. Oxford University Press. Eitan, Z. (2017). Cross-modal experience of musical pitch as space and motion: Current research and future challenges. In Body, sound and space in music and beyond: Multimodal explorations (pp. 49–68). Routledge.

234

References

Eitan, Z., & Granot, R. Y. (2006). How music moves: Musical parameters and listeners images of motion. Music Perception, 23(3), 221–248. Elleström, L. (2002). Divine madness: On interpreting literature, music, and the visual arts ironically. Bucknell University Press. Ellis, R., & Tucker, M. (2000). Micro-affordance: The potentiation of components of action by seen objects. British Journal of Psychology, 91(4), 451–471. Engel, A. K., Fries, P., König, P., Brecht, M., & Singer, W. (1999). Temporal binding, binocular rivalry, and consciousness. Consciousness and Cognition, 8(2), 128–151. Etzel, J. A., Johnsen, E. L., Dickerson, J., Tranel, D., & Adolphs, R. (2006). Cardiovascular and respiratory responses during musical mood induction. International Journal of Psychophysiology, 61(1), 57–69. Everett, Y. U. (2004). Parody with an ironic edge: Dramatic works by Kurt Weill, Peter Maxwell Davies, and Louis Andriessen. Music Theory Online, 10(4). Eysenck, M. W., & Keane, M. T. (2015). Cognitive psychology: A student’s handbook. Psychology Press. Fagg, A. H., & Arbib, M. A. (1998). Modeling parietal–premotor interactions in primate control of grasping. Neural Networks, 11(7–8), 1277–1303. Farah, M. J., & McClelland, J. L. (1992). Neural network models and cognitive neuropsychology. Psychiatric Annals, 22(3), 148–153. Fauconnier, G. (1985). Mental spaces: Aspects of meaning construction in natural language. MIT Press. Fauconnier, G., & Turner, M. (2002). The way we think: Conceptual blending and the mind’s hidden complexities. Basic Books. Feldman, J., & Narayanan, S. (2004). Embodied meaning in a neural theory of language. Brain and Language, 89(2), 385–392. Festinger, L., Gerard, H. B., Hymovitch, B., Kelley, H. H., & Raven, B. (1952). The influence process in the presence of extreme deviates. Human Relations, 5(4), 327–346. Forceville, C., & Jeulink, M. (2011). The flesh and blood of embodied understanding: The source-pathgoal schema in animation film. Pragmatics & Cognition, 19(1), 37–59. Fotheringhame, D. K., & Young, M. P. (1997). Neural coding schemes for sensory representation: Theoretical proposals and empirical evidence. Cognitive Neuroscience, 47–76. Foulds-Elliott, S. D., Thorpe, C. W., Cala, S. J., & Davis, P. J. (2000). Respiratory function in operatic singing: Effects of emotional connection. Logopedics Phoniatrics Vocology, 25(4), 151–168. Freedman, J. L., Birsky, J., & Cavoukian, A. (1980). Environmental determinants of behavioral contagion: Density and number. Basic and Applied Social Psychology, 1(2), 155–161. Friedmann, J. L. (2017). Music to climb by: Rising chromaticism in Max Steiner’s score for King Kong. Journal of Film Music, 10(2), 153–161. Fries, P. (2005). A mechanism for cognitive dynamics: Neuronal communication through neuronal coherence. Trends in Cognitive Sciences, 9(10), 474–480. Fujioka, T., Trainor, L. J., Large, E. W., & Ross, B. (2012). Internalized timing of isochronous sounds is represented in neuromagnetic beta oscillations. Journal of Neuroscience, 32(5), 1791–1802. Fuster, J. M., & Bressler, S. L. (2012). Cognit activation: A mechanism enabling temporal integration in working memory. Trends in Cognitive Sciences, 16(4), 207–218. Gabrielsson, A., & Juslin, P. N. (1996). Emotional expression in music performance: Between the performer’s intention and the listener’s experience. Psychology of Music, 24(1), 68–91. Galantucci, B., Fowler, C. A., & Turvey, M. T. (2006). The motor theory of speech reviewed. Psychonomic Bulletin & Review, 13, 361–377. Gallese, V. (2009). Mirror neurons, embodied simulation, and the neural basis of social identification. Psychoanalytic Dialogues, 19(5), 519–536. Gallese, V. (2010). Neuroscientific approach to intersubjectivity. In The embodied self: Dimensions, coherence, and disorders (pp. 77–92). Schattauer. Gallese, V. (2017). The empathic body in experimental aesthetics–embodied simulation and art. In Empathy (pp. 181–199). Palgrave Macmillan.

References

235

Gallese, V., Fadiga, L., Fogassi, L., & Rizzolatti, G. (1996). Action recognition in the premotor cortex. Brain, 119(2), 593–609. Garbarini, F., & Adenzato, M. (2004). At the root of embodied cognition: Cognitive science meets neurophysiology. Brain and Cognition, 56(1), 100–106. Gernsbacher, M. A., Keysar, B., Robertson, R. R., & Werner, N. K. (2001). The role of suppression and enhancement in understanding metaphors. Journal of Memory and Language, 45(3), 433–450. Gibbs Jr, R. W. (2001). Evaluating contemporary models of figurative language understanding. Metaphor and Symbol, 16(3–4), 317–333. Gibson, J. J. (1966). The senses considered as perceptual systems. Houghton-Mifflin. Gibson, J. J. (1979). The ecological approach to visual perception. Houghton-Mifflin. Gjerdingen, R. O. (1986). The formation and deformation of classic/romantic phrase schemata: A theoretical model and historical study. Music Theory Spectrum, 8, 25–43. Gjerdingen, R. O. (1994). Apparent motion in music? Music Perception, 11(4), 335–370. Gjerdingen, R. O., & Perrott, D. (2008). Scanning the dial: The rapid recognition of music genres. Journal of New Music Research, 37(2), 93–100. Godøy, R. I. (2003). Gestural imagery in the service of musical imagery. In Gesture-based communication in human-computer interaction: 5th international gesture workshop, GW 2003, Genova, Italy (pp. 55–62). Springer. Grady, J. (1997). Foundations of meaning: Primary metaphors and primary scenes. University of California: EScholarship. Grahn, J. A., & Brett, M. (2007). Rhythm and beat perception in motor areas of the brain. Journal of Cognitive Neuroscience, 19(5), 893–906. Grahn, J. A., & Rowe, J. B. (2009). Feeling the beat: Premotor and striatal interactions in musicians and nonmusicians during beat perception. Journal of Neuroscience, 29(23), 7540–7548. Grandjean, D., Sander, D., & Scherer, K. R. (2008). Conscious emotional experience emerges as a function of multilevel, appraisal-driven response synchronization. Consciousness and Cognition, 17(2), 484–495. Granot, R. Y., & Eitan, Z. (2011). Musical tension and the interaction of dynamic auditory parameters. Music Perception, 28(3), 219–246. Gregory, R. L. (1980). Perceptions as hypotheses. Philosophical Transactions of the Royal Society B: Biological Sciences, 290(1038), 181–197. Gridley, M. C., & Hoff, R. (2006). Do mirror neurons explain misattribution of emotions in music? Perceptual and Motor Skills, 102(2), 600–602. Grodal, T. (2009). Embodied visions: Evolution, emotion, culture, and film. Oxford University Press. Guenther, R. K. (1998). Human cognition. Prentice Hall. Guenther, R. K. (2002). Memory. In Foundations of cognitive psychology: Core readings. MIT Press. Hacohen, R., & Wagner, N. (1997). The communicative force of Wagner’s Leitmotifs: Complementary relationships between their connotations and denotations. Music Perception, 14(4), 445–475. Halfyard, J. K. (2004). Danny Elfman’s Batman: A film score guide. Scarecrow Press. Halfyard, J. K. (2013). Cue the big theme? The sound of the superhero. In The Oxford handbook of new audiovisual aesthetics (pp. 171–193). Oxford University Press. Hammerschmidt, K., & Jürgens, U. (2007). Acoustical correlates of affective prosody. Journal of Voice, 21(5), 531–540. Harris, R., & De Jong, B. M. (2014). Cerebral activations related to audition-driven performance imagery in professional musicians. PLoS One, 9(4), e93681. Hatten, R. S. (1994). Musical meaning in Beethoven: Markedness, correlation, and interpretation. Indiana University Press. Hatten, R. S. (2004). Interpreting musical gestures, topics, and tropes: Mozart, Beethoven, Schubert. Indiana University Press. Hauk, O., Johnsrude, I., & Pulvermüller, F. (2004). Somatotopic representation of action words in human motor and premotor cortex. Neuron, 41(2), 301–307.

236

References

Hecht, H., Kavelaars, J., Cheung, C. C., & Young, L. R. (2001). Orientation illusions and heart-rate changes during short-radius centrifugation. Journal of Vestibular Research, 11(2), 115–127. Heidemann, K. (2016). A system for describing vocal timbre in popular song. Music Theory Online, 22(1). Heine, E. (2018). Chromatic mediants and narrative context in film. Music Analysis, 37(1), 103–132. Heinichen, J. D. (1711). Neu erfundene und gründliche Anweisung wie ein Musik-liebender auf gewisse vortheilhafte Art könne zu vollkommener Erlernung des General-Basses. Schiller. Heldt, G. (2013). Music and levels of narration in film. Intellect. Herget, A. K. (2021). On music’s potential to convey meaning in film: A systematic review of empirical evidence. Psychology of Music, 49(1), 21–49. Hodhod, R. A., & Magerko, B. S. (2014). Computational creativity: Improv agents and conceptual blending. International Journal of Cognitive Informatics & Natural Intelligence, 8(2), 1–14. Hoeckner, B., Wyatt, E. W., Decety, J., & Nusbaum, H. (2011). Film music influences how viewers relate to movie characters. Psychology of Aesthetics, Creativity, and the Arts, 5(2), 146. Holmes, D. S. (1972). Repression or interference? A further investigation. Journal of Personality and Social Psychology, 22(2), 163. Hsu, H. C., & Su, L. I. (2014). Love in disguise: Incongruity between text and music in song. Journal of Pragmatics, 62, 136–150. Hulme, C., Maughan, S., & Brown, G. D. (1991). Memory for familiar and unfamiliar words: Evidence for a long-term memory contribution to short-term memory span. Journal of Memory and Language, 30(6), 685–701. Hunt, R. R. (2013). Precision in memory through distinctive processing. Current Directions in Psychological Science, 22(1), 10–15. Hurley, S. (2008). The shared circuits model (SCM): How control, mirroring, and simulation can enable imitation, deliberation, and mindreading. Behavioral and Brain Sciences, 31(1), 1–22. Huron, D., & Berec, J. (2009). Characterizing idiomatic organization in music: A theory and case study of musical affordances. Empirical Musicology Review, 4(3), 103–122. Huron, D., Kinney, D., & Precoda, K. (2006). Influence of pitch height on the perception of submissiveness and threat in musical passages. Empirical Musicology Review, 1(3), 170–177. Huron, D., & Margulis, E. H. (2010). Musical expectancy and thrills. In Handbook of music and emotion: Theory, research, applications. Oxford University Press. Huron, D. B. (2006). Sweet anticipation: Music and the psychology of expectation. MIT Press. Husain, G., Thompson, W. F., & Schellenberg, E. G. (2002). Effects of musical tempo and mode on arousal, mood, and spatial abilities. Music Perception, 20(2), 151–171. Hutcheon, L. (1985). A theory of parody: The teachings of twentieth-century art forms. Methuen. Iacoboni, M., Molnar-Szakacs, I., Gallese, V., Buccino, G., Mazziotta, J. C., & Rizzolatti, G. (2005). Grasping the intentions of others with one’s own mirror neuron system. PLoS Biology, 3(3), e79. Iacoboni, M., Woods, R. P., Brass, M., Bekkering, H., Mazziotta, J. C., & Rizzolatti, G. (1999). Cortical mechanisms of human imitation. Science, 286(5449), 2526–2528. Ibarretxe-Antuñano, I. (2013). The relationship between conceptual metaphor and culture. Intercultural Pragmatics, 10(2), 315–339. Jabbi, M., & Keysers, C. (2008). Inferior frontal gyrus activity triggers anterior insula response to emotional facial expressions. Emotion, 8(6), 775. Jacoby, L. L. (1978). On interpreting the effects of repetition: Solving a problem versus remembering a solution. Journal of Verbal Learning and Verbal Behavior, 17(6), 649–667. James, W. (1884). What is an emotion? Mind, 9(34). James, W. (1890). Principles of psychology (Vol. 2). Holt. Janata, P., & Grafton, S. T. (2003). Swinging in the brain: Shared neural substrates for behaviors related to sequencing and music. Nature Neuroscience, 6(7), 682–687. Johnson, M. (1987). The body in the mind: The bodily basis of meaning, imagination, and reason. University of Chicago Press.

References

237

Johnson, M. (2007). The meaning of the body. In W. Overton, U. Mueller, & J. Newman (Eds.), Developmental perspectives on embodiment and consciousness. Psychology Press. Johnson, M., & Lakoff, G. (2002). Why cognitive linguistics requires embodied realism. Cognitive Linguistics, 13(3), 245–264. Johnson, M., & Larson, S. (2003). ‘Something in the way she moves’—metaphors of musical motion. Metaphor and Symbol, 18, 63–84. Jones, M. R. (1993). Dynamics of musical patterns: How do melody and rhythm fit together. In Psychology and music: The understanding of melody and rhythm (pp. 67–92). Erlbaum Associates. Jones, M. R., & Boltz, M. (1989). Dynamic attending and responses to time. Psychological Review, 96, 459–491. Juslin, P. N. (2013). From everyday emotions to aesthetic emotions: Towards a unified theory of musical emotions. Physics of Life Reviews, 10(3), 235–266. Juslin, P. N., & Laukka, P. (2003). Communication of emotions in vocal expression and music performance: Different channels, same code? Psychological Bulletin, 129(5), 770. Juslin, P. N., Liljeström, S., Västfjäll, D., & Lundqvist, L. O. (2010). How does music evoke emotions? Exploring the underlying mechanisms. In P. N. Juslin & J. A. Sloboda (Eds.), Handbook of music and emotion: Theory, research, applications (pp. 605–642). Oxford University Press. Juslin, P. N., & Västfjäll, D. (2008). Emotional responses to music: The need to consider underlying mechanisms. Behavioral and Brain Sciences, 31(5), 559–575. Kendall, G. S. (2010). Meaning in electroacoustic music and the everyday mind. Organised Sound, 15(1), 63–74. Keyfitz, N. (1996). Review of Keeping together in time: Dance and drill in human history by McNeill, W. H. Contemporary Sociology, 25(3), 408–409. Keysers, C. (2011). The empathic brain: How the discovery of mirror neurons changes our understanding of human nature. Lulu.com. Khalfa, S., Roy, M., Rainville, P., Dalla Bella, S., & Peretz, I. (2008). Role of tempo entrainment in psychophysiological differentiation of happy and sad music? International Journal of Psychophysiology, 68(1), 17–26. Kinsbourne, M. (2005). Imitation as entrainment: Brain mechanisms and social consequences. Perspectives on Imitation: From Neuroscience to Social Science, 2, 163–172. Kintsch, W., Patel, V. L., & Ericsson, K. A. (1999). The role of long-term working memory in text comprehension. Psychologia, 42(4), 186–198. Koelsch, S., Fritz, T., Cramon, D. Y. V., Müller, K., & Friederici, A. D. (2006). Investigating emotion with music: An fMRI study. Human Brain Mapping, 27(3), 239–250. Koestler, A. (1964). The act of creation. Macmillan. Kohler, E., Keysers, C., Umilta, M. A., Fogassi, L., Gallese, V., & Rizzolatti, G. (2002). Hearing sounds, understanding actions: Action representation in mirror neurons. Science, 297(5582), 846–848. Köhler, W. (1929). Gestalt psychology. H. Liveright. Kostka, V. (2016). Linda Hutcheon’s theory of parody and its application to postmodern music. Avant: Pismo Awangardy Filozoficzno-Naukowej, (1), 67–73. Kövecses, Z. (2002). Metaphor: A practical introduction. Oxford University Press. Krueger, J. W. (2011). Doing things with music. Phenomenology and the Cognitive Sciences, 10(1), 1–22. Krumhansl, C. L. (1991). Music psychology: Tonal structures in perception and memory. Annual Review of Psychology, 42(1), 277–303. Krumhansl, C. L. (1998). Topic in music: An empirical study of memorability, openness, and emotion in Mozart’s String Quintet in C Major and Beethoven’s String Quartet in A Minor. Music Perception, 16(1), 119–134. Kruschke, J. K. (2008). Models of categorization. In The Cambridge handbook of computational psychology (pp. 267–301). Cambridge University Press.

238

References

Kühn, S., Werner, A., Lindenberger, U., & Verrel, J. (2014). Acute immobilisation facilitates premotor preparatory activity for the non-restrained hand when facing grasp affordances. NeuroImage, 92, 69–73. Küssner, M. B., Tidhar, D., Prior, H. M., & Leech-Wilkinson, D. (2014). Musicians are more consistent: Gestural cross-modal mappings of pitch, loudness and tempo in real-time. Frontiers in Psychology, 5, 789. Kvavilashvili, L., & Mandler, G. (2004). Out of one’s mind: A study of involuntary semantic memories. Cognitive Psychology, 48(1), 47–94. Laird, J. D. (2007). Feelings: The perception of self. Oxford University Press. Lakin, J. L., Jefferis, V. E., Cheng, C. M., & Chartrand, T. L. (2003). The chameleon effect as social glue: Evidence for the evolutionary significance of nonconscious mimicry. Journal of Nonverbal Behavior, 27(3), 145–162. Lakoff, G. (1987). Women, fire, and dangerous things: What categories reveal about the mind. University of Chicago Press. Lakoff, G. (2014). Mapping the brain’s metaphor circuitry: Metaphorical thought in everyday reason. Frontiers in Human Neuroscience, 8, 958. Lakoff, G., & Johnson, M. (1980). Metaphors we live by. University of Chicago Press. Lakoff, G., & Sweetser, E. (1994). Foreword to Gilles Fauconnier, mental spaces. In Mental spaces (pp. 1–7). Cambridge University Press. Langacker, R. W. (1987). Foundations of cognitive grammar: Theoretical prerequisites (Vol. 1). Stanford University Press. Large, E. W., & Palmer, C. (2002). Perceiving temporal regularity in music. Cognitive Science, 26(1), 1–37. Large, E. W., & Snyder, J. S. (2009). Pulse and meter as neural resonance. Annals of the New York Academy of Sciences, 1169(1), 46–57. Larson, S. (2012). Musical forces: Motion, metaphor, and meaning in music. Indiana University Press. Lastinger, D. L. (2011). The effect of background music on the perception of personality and demographics. Journal of Music Therapy, 48(2), 208–225. Lee, Y. T., & Tsai, S. J. (2010). The mirror neuron system may play a role in the pathogenesis of mass hysteria. Medical Hypotheses, 74(2), 244–245. Lehman, F. (2013). Hollywood cadences: Music and the structure of cinematic expectation. Music Theory Online, 19(4). Lehman, F. (2018). Hollywood harmony: Musical wonder and the sound of cinema. Oxford University Press. Leinberger, C. (2004). Ennio Morricone’s the good, the bad and the ugly: A film score guide. Scarecrow Press. Leman, M., Moelants, D., Varewyck, M., Styns, F., van Noorden, L., & Martens, J. P. (2013). Activating and relaxing music entrains the speed of beat synchronized walking. PLoS One, 8(7), e67932. Levenson, R. W., Ekman, P., & Friesen, W. V. (1990). Voluntary facial action generates emotion-specific autonomic nervous system activity. Psychophysiology, 27(4), 363–384. Lévêque, Y., Muggleton, N., Stewart, L., & Schön, D. (2013). Involvement of the larynx motor area in singing-voice perception: A TMS study. Frontiers in Psychology, 4, 418. Lévêque, Y., & Schon, D. (2015). Modulation of the motor cortex during singing-voice perception. Neuropsychologia, 70, 58–63. Levitin, D. J. (2006). This is your brain on music: The science of a human obsession. Penguin. Lidov, D. (2005). Is language a music?: Writings on musical form and signification. Indiana University Press. Lima, C. F., Krishnan, S., & Scott, S. K. (2016). Roles of supplementary motor areas in auditory processing and auditory imagery. Trends in Neurosciences, 39(8), 527–542. Lipps, T. (1903). Ästhetik. Leopold Voss Verlag. Lipscomb, S. D. (1995). Cognition of musical and visual accent structure alignment in film and animation [Unpublished PhD dissertation, University of California, Los Angeles].

References

239

Lipscomb, S. D., & Kendall, R. A. (1994). Perceptual judgement of the relationship between musical and visual components in film. Psychomusicology: A Journal of Research in Music Cognition, 13 (1–2), 60. London, J. (2012). Hearing in time: Psychological aspects of musical meter. Oxford University Press. Lorenzi-Filho, G., Dajani, H. R., Leung, R. S., Floras, J. S., & Bradley, T. D. (1999). Entrainment of blood pressure and heart rate oscillations by periodic breathing. American Journal of Respiratory and Critical Care Medicine, 159(4), 1147–1154. Love, S. C. (2017). An ecological description of jazz improvisation. Psychomusicology: Music, Mind, and Brain, 27(1), 31. Macken, W. J., Tremblay, S., Houghton, R. J., Nicholls, A. P., & Jones, D. M. (2003). Does auditory streaming require attention? Evidence from attentional selectivity in short-term memory. Journal of Experimental Psychology: Human Perception and Performance, 29(1), 43. Makris, S., Hadar, A. A., & Yarrow, K. (2013). Are object affordances fully automatic? A case of covert attention. Behavioral Neuroscience, 127(5), 797. Marshall, S. K., & Cohen, A. J. (1988). Effects of musical soundtracks on attitudes toward animated geometric figures. Music Perception, 6(1), 95–112. Masataka, N. (2006). Preference for consonance over dissonance by hearing newborns of deaf parents and of hearing parents. Developmental Science, 9(1), 46–50. Mason, F. (2011). Hollywood’s detectives: Crime series in the 1930s and 1940s from the Whodunnit to hard-boiled Noir. Springer. Mathy, F., & Feldman, J. (2012). What’s magic about magic numbers? Chunking and data compression in short-term memory. Cognition, 122(3), 346–362. Mattheson, J. (1739). Der vollkommene Capellmeister. Herold. Mayville, J. M., Jantzen, K. J., Fuchs, A., Steinberg, F. L., & Kelso, J. S. (2002). Cortical and subcortical networks underlying syncopated and synchronized coordination revealed using fMRI. Human Brain Mapping, 17(4), 214–229. McAngus Todd, N. P., O’Boyle, D. J., & Lee, C. S. (1999). A sensory-motor theory of rhythm, time perception and beat induction. Journal of New Music Research, 28(1), 5–28. McClelland, J. L. (2000). Connectionist models of memory. In The Oxford handbook of memory (pp. 583–596). Oxford University Press. McGettigan, C., Walsh, E., Jessop, R., Agnew, Z. K., Sauter, D. A., Warren, J. E., & Scott, S. K. (2015). Individual differences in laughter perception reveal roles for mentalizing and sensorimotor systems in the evaluation of emotional authenticity. Cerebral Cortex, 25(1), 246–257. McLeod, K. (2006). “A fifth of Beethoven”: Disco, classical music, and the politics of inclusion. American Music, 347–363. McNeill, W. H. (1997). Keeping together in time: Dance and drill in human history. Harvard University Press. Medin, D. L., & Coley, J. D. (1998). Concepts and categorization. In Perception and cognition at century’s end: Handbook of perception and cognition (pp. 403–439). Elsevier Science. Menon, V., & Levitin, D. J. (2005). The rewards of music listening: Response and physiological connectivity of the mesolimbic system. Neuroimage, 28(1), 175–184. Merker, B. H., Madison, G. S., & Eckerdal, P. (2009). On the role and origin of isochrony in human rhythmic entrainment. Cortex, 45(1), 4–17. Merleau-Ponty, M., & Smith, C. (1962). Phenomenology of perception (Vol. 26). Routledge. Meyer-Kalkus, R. (2007). Work, rhythm, dance. Embodiment in Cognition and Culture, 71, 165–181. Milgram, S., Bickman, L., & Berkowitz, L. (1969). Note on the drawing power of crowds of different size. Journal of Personality and Social Psychology, 13(2), 79. Miller, C. T., & Cohen, Y. E. (2010). Vocalizations as auditory objects: Behavior and neurophysiology. In Primate neuroethology. Oxford University Press. Millet, B., Chattah, J., & Ahn, S. (2021). Soundtrack design: The impact of music on visual attention and affective responses. Applied Ergonomics, 93(103301), 1–9.

240

References

Molenberghs, P., Cunnington, R., & Mattingley, J. B. (2012). Brain regions with mirror properties: A meta-analysis of 125 human fMRI studies. Neuroscience & Biobehavioral Reviews, 36(1), 341–349. Molnar-Szakacs, I., & Overy, K. (2006). Music and mirror neurons: From motion to ‘e’motion. Social Cognitive and Affective Neuroscience, 1(3), 235–241. Morris, P. (1988). Memory research: Past mistakes and future prospects. In Growth points in cognition (pp. 91–110). Routledge. Motazedian, T. (2016). To key or not to key: Tonal design in film music [Doctoral dissertation, Yale University]. Mukamel, R., Ekstrom, A. D., Kaplan, J., Iacoboni, M., & Fried, I. (2010). Single-neuron responses in humans during execution and observation of actions. Current Biology, 20(8), 750–756. Murphy, S. (2006). The major tritone progression in recent Hollywood science fiction films. Music Theory Online, 12(2). Murphy, S. (2022). Three audiovisual correspondences in the main title for Vertigo. Music Theory Online, 28(1). Murphy, S. T., & Zajonc, R. B. (1993). Affect, cognition, and awareness: Affective priming with optimal and suboptimal stimulus exposures. Journal of Personality and Social Psychology, 64(5), 723. Nairne, J. S. (2002). The myth of the encoding-retrieval match. Memory, 10(5–6), 389–395. Neumann, R., & Strack, F. (2000). “Mood contagion”: The automatic transfer of mood between persons. Journal of Personality and Social Psychology, 79(2), 211. Norman, J. (2002). Two visual systems and two theories of perception: An attempt to reconcile the constructivist and ecological approaches. Behavioral and Brain Sciences, 25(1), 73–96. Nosal, A. P., Keenan, E. A., Hastings, P. A., & Gneezy, A. (2016). The effect of background music in shark documentaries on viewers’ perceptions of sharks. PLoS One, 11(8), e0159279. Nozaradan, S., Peretz, I., Missal, M., & Mouraux, A. (2011). Tagging the neuronal entrainment to beat and meter. Journal of Neuroscience, 31(28), 10234–10240. Oden, C. (2023). “We are dancing, we are flying”: The feeling of flight in dance scenes from recent popular film. Society for Music Theory: Videocast Journal (SMT-V), 9(1). Oman, C. M. (1998). Sensory conflict theory and space sickness: Our changing perspective. Journal of Vestibular Research, 8(1), 51–56. Oman, C. M. (2003). Neurovestibular adaptation to spaceflight: Research progress. Journal of Vestibular Research, 12(5–6), 201–203. Ortony, A. (1993). Metaphor and thought (2nd ed.). Cambridge University Press. Osiurak, F., Rossetti, Y., & Badets, A. (2017). What is an affordance? 40 years later. Neuroscience & Biobehavioral Reviews, 77, 403–417. Overy, K., & Molnar-Szakacs, I. (2009). Being together in time: Musical experience and the mirror neuron system. Music Perception, 26(5), 489–504. Owren, M. J., & Bachorowski, J. A. (2007). Measuring emotion-related vocal acoustics. In Handbook of emotion elicitation and assessment (pp. 239–266). Oxford University Press. Ozturk, O., Krehm, M., & Vouloumanos, A. (2013). Sound symbolism in infancy: Evidence for soundshape cross-modal correspondences in 4-month-olds. Journal of Experimental Child Psychology, 114(2), 173–186. Patel, A. D., Iversen, J. R., Chen, Y., & Repp, B. H. (2005). The influence of metricality and modality on synchronization with a beat. Experimental Brain Research, 163(2), 226–238. Phillips, W. A., & Singer, W. (1997). In search of common foundations for cortical computation. The Behavioral and Brain Sciences, 20(4), 657–683. Phillips-Silver, J., Aktipis, C. A., & Bryant, G. (2010). The ecology of entrainment: Foundations of coordinated rhythmic movement. Music Perception, 28(1), 3–14. Pinna, S. (2017). Cognition as organism-environment interaction. In Extended cognition and the dynamics of algorithmic skills (pp. 19–37). Springer. Plantinga, J., & Trehub, S. E. (2014). Revisiting the innate preference for consonance. Journal of Experimental Psychology. Human Perception and Performance, 40(1), 40–49.

References

241

Platel, H., Price, C., Baron, J. C., Wise, R., Lambert, J., Frackowiak, R. S., Lechevalier, B., & Eustache, F. (1997). The structural components of music perception. A functional anatomical study. Brain: A Journal of Neurology, 120(2), 229–243. Ploog, D. W. (1992). The evolution of vocal communication. In H. Papoušek, U. Jürgens, & M. Papoušek (Eds.), Nonverbal vocal communication: Comparative and developmental approaches (pp. 6–30). Cambridge University Press; Editions de la Maison des Sciences de l’Homme. Proverbio, A. M., Azzari, R., & Adorni, R. (2013). Is there a left hemispheric asymmetry for tool affordance processing? Neuropsychologia, 51(13), 2690–2701. Ramachandran, V. S., & Hubbard, E. M. (2001). Synaesthesia—a window into perception, thought and language. Journal of Consciousness Studies, 8(12), 3–34. Rapée, E. (1925). Encyclopaedia of music for pictures. Belwin. Ratner, L. G. (1980). Classic music: Expression, form, and style. Schirmer Books; Collier Macmillan Publishers. Reason, J. T., & Brand, J. J. (1975). Motion sickness. Academic Press. Repp, B. H., & Penel, A. (2004). Rhythmic movement is attracted more strongly to auditory than to visual rhythms. Psychological Research, 68(4), 252–270. Reybrouck, M. (2005). A biosemiotic and ecological approach to music cognition: Event perception between auditory listening and cognitive economy. Axiomathes, 15(2), 229–266. Reybrouck, M. (2010). Music cognition and real-time listening: Denotation, cue abstraction, route description and cognitive maps. Musicae Scientiae, 14(2_suppl), 187–202. Reybrouck, M. (2012). Musical sense-making and the concept of affordance: An ecosemiotic and experiential approach. Biosemiotics, 5(3), 391–409. Ritchie, L. D. (2004). Lost in “conceptual space”: Metaphors of conceptual integration. Metaphor and Symbol, 19(1), 31–50. Rizzolatti, G., & Fadiga, L. (1998). Grasping objects and grasping action meanings: The dual role of monkey rostroventral premotor cortex (area F5). Sensory Guidance of Movement, 218, 81–103. Rizzolatti, G., Fadiga, L., Gallese, V., & Fogassi, L. (1996). Premotor cortex and the recognition of motor actions. Cognitive Brain Research, 3(2), 131–141. Rochat, M. J., Caruana, F., Jezzini, A., Escola, L., Intskirveli, I., Grammont, F., Gallese, V., Rizzolatti, G. and Umiltà, M. A. (2010). Responses of mirror neurons in area F5 to hand and tool grasping observation. Experimental Brain Research, 204(4), 605–616. Rosch, E. (1978). Principles of categorization. In Cognition and categorization (pp. 27–48). Erlbaum Associates. Rosch, E., & Mervis, C. B. (1975). Family resemblances: Studies in the internal structure of categories. Cognitive Psychology, 7(4), 573–605. Rötter, G. (1994). The perception of form and its relation to emotional listening and physiological data. In I. Deliège (Ed.), Proceedings of the 3rd international conference on music perception and cognition (pp. 277–278). European Society for the Cognitive Sciences of Music. Rusconi, E., Kwan, B., Giordano, B. L., Umiltà, C., & Butterworth, B. (2006). Spatial representation of pitch height: The SMARC effect. Cognition, 99(2), 113–129. Ryan, J. D., Althoff, R. R., Whitlow, S., & Cohen, N. J. (2000). Amnesia is a deficit in relational memory. Psychological Science, 11(6), 454–461. Sachs, C., & Kunst, J. (1962). The wellsprings of music. M. Nijhoff. Sakreida, K., Effnert, I., Thill, S., Menz, M. M., Jirak, D., Eickhoff, C. R., Ziemke, T., Eickhoff, S. B., Borghi, A. M., & Binkofski, F. (2016). Affordance processing in segregated parieto-frontal dorsal stream sub-pathways. Neuroscience & Biobehavioral Reviews, 69, 89–112. Saslaw, J. (1996). Forces, containers, and paths: The role of body-derived image schemas in the conceptualization of music. Journal of Music Theory, 40(2), 217–243. Sato, T. G., Ohsuga, M., & Moriya, T. (2012). Increase in the timing coincidence of a respiration event induced by listening repeatedly to the same music track. Acoustical Science and Technology, 33(4), 255–261. Sayrs, E. (2003). Narrative, metaphor, and conceptual blending in ‘the hanging tree.’ Music Theory Online, 9(1).

242

References

Schacter, D. L. (1987). Implicit memory: History and current status. Journal of Experimental Psychology: Learning, Memory, and Cognition, 13(3), 501. Schacter, D. L., & Tulving, E. (1994). What are the memory systems of 1994? In D. L. Schacter & E. Tulving (Eds.), Memory systems 1994 (pp. 1–38). The MIT Press. Schank, R. C., & Abelson, R. P. (1975). Scripts, plans, and knowledge. IJCAI, 75, 151–157. Schelle, M. (1999). The score: Interviews with film composers. Silman-James Press. Scherer, K. R., Clark-Polner, E., & Mortillaro, M. (2011). In the eye of the beholder? Universality and cultural specificity in the expression and perception of emotion. International Journal of Psychology, 46(6), 401–435. Scherer, K. R., & Coutinho, E. (2013). How music creates emotion: A multifactorial process approach. In The emotional power of music: Multidisciplinary perspectives on musical arousal, expression, and social control (pp. 121–145). Oxford University Press. Scherer, K. R., Trznadel, S., Fantini, B., & Sundberg, J. (2017). Recognizing emotions in the singing voice. Psychomusicology: Music, Mind, and Brain, 27(4), 244. Schiavetto, A., Cortese, F., & Alain, C. (1999). Global and local processing of musical sequences: An event-related brain potential study. Neuroreport, 10(12), 2467–2472. Schmidt, L. (2010). A popular avant-garde: The paradoxical tradition of electronic and atonal sounds in sci-fi music scoring. In Sounds of the future: Essays on music in science fiction films. McFarland & Co. Shepard, R. N. (2009). One cognitive psychologist’s quest for the structural grounds of music cognition. Psychomusicology: Music, Mind and Brain, 20(1–2), 130. Shevy, M. (2008). Music genre as cognitive schema: Extramusical associations with country and hiphop music. Psychology of Music, 36(4), 477–498. Singer, T., Seymour, B., O’Doherty, J., Kaube, H., Dolan, R. J., & Frith, C. D. (2004). Empathy for pain involves the affective but not sensory components of pain. Science, 303(5661), 1157–1162. Sirius, G., & Clarke, E. F. (1994). The perception of audiovisual relationships: A preliminary study. Psychomusicology, 13(1–2), 119. Smith, J. D., Zakrzewski, A. C., Johnson, J. M., Valleau, J. C., & Church, B. A. (2016). Categorization: The view from animal cognition. Behavioral Sciences, 6(2), 12. Spiers, H. J., Maguire, E. A., & Burgess, N. (2001). Hippocampal amnesia. Neurocase, 7(5), 357–382. Spitzer, M. (2004). Metaphor and musical thought. Chicago University Press. Spitzer, M. (2018). Conceptual blending and musical emotion. Musicae Scientiae, 22(1), 24–37. Strack, F., Martin, L. L., & Stepper, S. (1988). Inhibiting and facilitating conditions of the human smile: A nonobtrusive test of the facial feedback hypothesis. Journal of Personality and Social Psychology, 54(5), 768. Sundberg, J., Patel, S., Bjorkner, E., & Scherer, K. R. (2011). Interdependencies among voice source parameters in emotional speech. IEEE Transactions on Affective Computing, 2(3), 162–174. Sweeny, T., Guzman-Martinez, E., Ortega, L., Grabowecky, M., & Suzuki, S. (2012). Sounds exaggerate visual shape. Cognition, 124(2), 194–200. Tagg, P. (1989). An anthropology of stereotypes in TV music? Swedish Musicological Journal, 71, 19–42. Tagg, P. (1999). Introductory notes to the semiotics of music, version 3. Retrieved December 10, 2020, from www.tagg.org/xpdfs/semiotug.pdf Tagg, P. (2012). Music’s meanings: A modern musicology for non-musos. The Mass Media Music Scholars’ Press. Tagg, P., & Clarida, B. (2003). Ten little title tunes. Mass Media Scholars’ Press. Talmy, L. (1983). How language structures space. In H. L. Pick, Jr. & L. P. Acredolo (Eds.), Spatial orientation: Theory, research, and application. Plenum Press. Tan, S. L., Baxa, J., & Spackman, M. P. (2010). Effects of built-in audio versus unrelated background music on performance in an adventure role-playing game. International Journal of Gaming and Computer–Mediated Simulations (IJGCMS), 2(3), 1–23.

References

243

Tan, S. L., Spackman, M. P., & Wakefield, E. M. (2008). Effects of diegetic and non–diegetic presentation of film music on viewers’ interpretation of film narrative. In Conference proceedings for the 2008 international conference of music perception and cognition (pp. 588–593). Department of Psychology, Hokkaido University. Thagard, P., & Aubie, B. (2008). Emotional consciousness: A neural model of how cognitive appraisal and somatic perception interact to produce qualitative experience. Consciousness and Cognition, 17(3), 811–834. Thagard, P., & Stewart, T. C. (2011). The AHA! experience: Creativity through emergent binding in neural networks. Cognitive Science, 35(1), 1–33. Thompson, W. F., Russo, F. A., & Sinclair, D. (1994). Effects of underscoring on the perception of closure in filmed events. Psychomusicology: A Journal of Research in Music Cognition, 13(1–2), 9. Tischler, M., & Morey-Holton, E. (1992). Space research on organs and tissues. In Space programs and technologies conference (p. 1345). American Institute of Aeronautics and Astronautics. Töpper, J., & Schwan, S. (2008). James Bond in angst?: Inferences about protagonists’ emotional states in films. Journal of Media Psychology: Theories, Methods, and Applications, 20(4), 131. Trainor, L. J., Tsang, C. D., & Cheung, V. H. (2002). Preference for sensory consonance in 2-and 4-month-old infants. Music Perception, 20(2), 187–194. Trost, W., & Vuilleumier, P. (2013). Rhythmic entrainment as a mechanism for emotion induction by music: A neurophysiological perspective. In The emotional power of music: Multidisciplinary perspectives on musical arousal, expression, and social control (pp. 213–225). Oxford University Press. Tsougras, C., & Stefanou, D. (2018). Embedded blends and meaning construction in Modest Mussorgsky’s Pictures at an Exhibition. Musicae Scientiae, 22(1), 38–56. Tucker, M., & Ellis, R. (1998). On the relations between seen objects and components of potential actions. Journal of Experimental Psychology: Human Perception and Performance, 24(3), 830–846. Tulving, E. (1983). Elements of episodic memory. Oxford University Press. Vandeloise, C. (2007). A taxonomy of basic natural entities. In M. Aurnague, M. Hickmann, & L. Vieu (Eds.), Categorization of spatial entities in language and cognition. John Benjamins Publishing Company. Velasco, C., Woods, A., Petit, O., Cheok, A., & Spence, C. (2016). Crossmodal correspondences between taste and shape, and their implications for product packaging: A review. Food Quality and Preference, 52, 17–26. van der Velde, F., Forth, J., Nazareth, D. S., & Wiggins, G. A. (2017). Linking neural and symbolic representation and processing of conceptual structures. Frontiers in Psychology, 8, 1297. Wagner, A. D., Stebbins, G. T., Masciari, F., Fleischman, D. A., & Gabrieli, J. D. (1998). Neuropsychological dissociation between recognition familiarity and perceptual priming in visual long-term memory. Cortex, 34(4), 493–511. Wagner, S., Winner, E., Cicchetti, D., & Gardner, H. (1981). “Metaphorical” mapping in human infants. Child Development, 728–731. Walker, M. (1992). Film Noir: Introduction. In The movie book of film Noir (pp. 8–38). Studio Vista. Ward, J. (2015). The student’s guide to cognitive neuroscience. Psychology Press. Warren, J. E., Sauter, D. A., Eisner, F., Wiland, J., Dresner, M. A., Wise, R. J., Rosen, S. and Scott, S. K. (2006). Positive emotions preferentially engage an auditory-motor “mirror” system. Journal of Neuroscience, 26(50), 13067–13075. Warren, W. H. (1984). Perceiving affordances: Visual guidance of stair climbing. Journal of Experimental Psychology: Human Perception and Performance, 10(5), 683–703. Waters, R. H., & Leeper, R. (1936). The relation of affective tone to the retention of experiences of daily life. Journal of Experimental Psychology, 19(2), 203. Watts, C. R., & Hall, M. D. (2008). Timbral influences on vocal pitch-matching accuracy. Logopedics Phoniatrics Vocology, 33(2), 74–82. Welch, R. (1999). Meaning, attention, and the “unity assumption” in the intersensory bias of spatial and temporal perceptions. Advances in Psychology, 129, 371–387.

244

References

Wheeler, L. (1966). Toward a theory of behavioral contagion. Psychological Review, 73(2), 179. Wicker, B., Keysers, C., Plailly, J., Royet, J., Gallese, V., & Rizzolatti, G. (2003). Both of us disgusted in my insula: The common neural basis of seeing and feeling disgust. Neuron (Cambridge, Mass.), 40(3), 655–664. Widmann, A., Kujala, T., Tervaniemi, M., Kujala, A., & Schröger, E. (2004). From symbols to sounds: Visual symbolic information activates sound representations. Psychophysiology, 41(5), 709–715. Wiggins, G. A. (2010). Cue abstraction, paradigmatic analysis and information dynamics: Towards music analysis by cognitive model. Musicae Scientiae, 14(2_suppl), 307–331. Will, U., & Berg, E. (2007). Brain wave synchronization and entrainment to periodic acoustic stimuli. Neuroscience Letters, 424(1), 55–60. Wilson, S. M., Saygin, A. P., Sereno, M. I., & Iacoboni, M. (2004). Listening to speech activates motor areas involved in speech production. Nature Neuroscience, 7(7), 701–702. Wingstedt, J., Brändström, S., & Berg, J. (2008). Young adolescents’ usage of narrative functions of media music by manipulation of musical expression. Psychology of Music, 36(2), 193–214. Winkler, I., & Cowan, N. (2005). From sensory to long-term memory: Evidence from auditory memory reactivation studies. Experimental Psychology, 52(1), 3–20. Winold, C. A. (1963). The effects of changes in harmonic tension upon listener response. Indiana University. Winters, B. (2007). Erich Wolfgang Korngold’s the adventures of Robin Hood: A film score guide (Vol. 6). Scarecrow Press. Wittgenstein, L. (1953). Philosophische Untersuchungen. Kegan Paul. Wühr, P., & Müsseler, J. (2002). Blindness to response-compatible stimuli in the psychological refractory period paradigm. Visual Cognition, 9(4–5), 421–457. Young, M. D. (2013). Musical topics in the comic book superhero film genre [PhD dissertation, The University of Texas at Austin]. Zanto, T. P., Snyder, J. S., & Large, E. W. (2006). Neural correlates of rhythmic expectancy. Advances in Cognitive Psychology, 2(2), 221. Zatorre, R. J., Chen, J. L., & Penhune, V. B. (2007). When the brain plays music: Auditory-motor interactions in music perception and production. Nature Reviews Neuroscience, 8(7), 547–558. Zatorre, R. J., Mondor, T. A., & Evans, A. C. (1999). Auditory attention to space and frequency activates similar cerebral systems. Neuroimage, 10(5), 544–554. Zbikowski, L. M. (1998). Metaphor and music theory: Reflections from cognitive science. Music Theory Online, 4(1). Zbikowski, L. M. (1999). The blossoms of ‘Trockne Blumen’: Music and text in the early nineteenth century. Music Analysis, 18(3), 307–345. Zbikowski, L. M. (2002). Conceptualizing music: Cognitive structure, theory, and analysis. Oxford University Press. Zbikowski, L. M. (2007). Aspects of meaning construction in music: Toward a cognitive grammar of music. Almen Semiotik, 17, 43–72. Zbikowski, L. M. (2009). Music, language, and multimodal metaphor. Multimodal Metaphor, 359–381. Zentner, M. R., & Kagan, J. (1998). Infants’ perception of consonance and dissonance in music. Infant Behavior and Development, 21(3), 483–492. van der Zwaag, M. D., Westerink, J. H., & van den Broek, E. L. (2011). Emotional and psychophysiological responses to tempo, mode, and percussiveness. Musicae Scientiae, 15(2), 250–269.

FILM, FILMMAKER, COMPOSER INDEX

15 Minutes 12–14, 56–57, 103–104 1917 223 2001: A Space Odyssey 73, 79, 84n26, 156 2010: The Year We Made Contact 73 300 69–70 45 Years 223–224 A.I. 224 About Time 114n3 Adventures of Robin Hood The 147–150, 163n2 Africa: The Serengeti 17 Airplane! 135, 136, 142 Alexander Nevsky 27 Alice in Wonderland (1951) 74 Alice in Wonderland (2010) 40–41 Alkan, Charles: Funeral March on the Death of a Par 97n6 Amadeus 58, 63n15 American Animals 10 American Beauty 62n4 An Education 71–72 August Rush 58 Avatar 45 Avengers Endgame 224 Back to the Future 76, 114n3 Barbarian 19n9 Batman 64, 67, 76–77, 81, 97n7, 112–113, 115n16, 115n17 Batman Forever 115n15 Batman TV series (1960s) 109–110 Beast 224–225 Beethoven, Ludwig V.: 211n6; Eroica Symphony 114n4; Piano Sonata #12 97n6; Symphony #3 97n6; Symphony #9 185 Being There 154–156

Berlioz, Hector: Marche Funèbre Pour la Dernière Scène 97n6 Big Night 156–157 Blade Runner 100 The Blair Witch Project 99n29 Bowling for Columbine 133, 134 Byrne, David 102 Cabaret 45–46 Captain America 111–112 Captain Blood 108–109, 115n11 Cassandra’s Dream 225 Charlie and the Chocolate Factory (2005) 41–43 Chicken Run 225 Chopin, Frédéric: Piano Sonata #2 in B♭-minor 97n6, 114n4 Citizen Kane 50, 59–62 City Lights 28 Cloudy with a Chance of Meatballs 83n24 Con Air 128, 129 The Conversation 26, 34, 44, 46–47, 57, 116, 124 Copland, Aaron 62n4 Corigliano, John 151–153, 163n4 Count Basie 110; “Blues in Hoss’ Flat” 29 Crouching Tiger, Hidden Dragon 42 Dalí, Salvador 75 Dance of the 41 160 Dancer in the Dark 33 Dangerous Liaisons 100–101 Davis, Don 78, 89 Deadpool 225–226 Donizetti, Gaetano: “Il Dolce Suono” (Lucia di Lammermoor) 55, 63n7 Dr. No 97n5, 110

246

FILM, FILMMAKER, COMPOSER INDEX

Edward Scissorhands 10–11 Elfman, Danny 77, 112, 115n17 Englishby, Paul 72 Eric Serra: “Diva Dance” 55 The Errand Boy 28 Face Off 136–137, 138 Fantasia 164n9 Farewell My Lovely 25–26, 82n12 The Father 32–33 Fenton, George 101 The Fifth Element 55 Film Stars Don’t Die In Liverpool 51 Finding Neverland 59, 66–67 Fools Rush In 118 Fox, Charles 111

Leone, Sergio 106 The Lives of Others 122–123 The Lobster 227 Looper 114n3 Lord of the Rings franchise 98n16 Lord of the Rings: Return of The King 9 Lord of War 121–122

Hefti, Neal 110 He Loves Me, He Loves Me Not 161–163 Herrmann, Bernard 60–62, 88 The Hindenburg 38–39 Home Alone 100, 102–103 Hot Tub Time Machine 114n3 The Hours 29, 53 The Hunger Games: Mockingjay – Part I 227 Hyams, Peter 73

Mahler, Gustav 144n9; Symphony #5 97n6 Maleficent 86–87 Mancina, Mark 59 Mancini, Henry 110 The Mandalorian 107 Marinelli, Anthony 56–57, 103 The Mask of Zorro 100 The Matrix 69, 78, 88–89 Meet the Parents 105–106, 137, 139 Mendelssohn, Felix: Song Without Words Op. 62, #3 114n4 Midnight Express 8 Midsommar 226 Miles Davis: “So What” 104 The Milk of Sorrow 31 Minority Report 120–121 Mission Impossible 67–68, 226 Mission Impossible II 29–30 Monty Norman 110 Morricone, Ennio 106, 114n7 Moulin Rouge! 51–52 Mozart, Wolfgang A. 58, 144n9, 211n6; Piano Sonata K. 332 204–205 M Squad 110 Mussorgsky, Modest: “Night on Bald Mountain” 159, 164n9

The Iguanas: “Para Dónde Vas?” 118 Imani Coppola: “I’m a Tree” 129 Inception 31, 36n16 The Incredibles 41–42 Interview with the Vampire 39

Nat King Cole: “L-O-V-E” 117, 161 Newman, Randy 105 Newman, Thomas 62n4 Newton Howard, James 86–87 Nickelback: “Hero” 115n15

James Bond franchise 110 Jaws 85, 92–96, 99n26, 99n28, 99n29, 135 Jumper 114n3 Jurassic Park 39–40 Jurassic World: Fallen Kingdom 14–15

Oldfield, Mike 68

Gladiator 12, 16 The Good, the Bad and the Ugly 106–107, 114n7 The Grand Budapest Hotel 139–141 Gravity 8–9, 19n11 Gregson-Williams, Rupert 113 Grieg, Edward: Funeral March 97n6 Guess Who 132, 133

Kaczmarek, Jan A. P. 59 King Kong 38, 48n4 Korngold, Erich W.: 108, 147, 163n2; Rosen aus Florida 163n3 Kubrick, Stanley 73, 84n26 Kung Fu Panda 91 Lady and the Tramp 97n5 Lars and the Real Girl 116–117 The Last Emperor 102

Paul McCartney: “Ebony and Ivory” 132 Peter Gunn 110 The Pink Panther 110 Pirates of the Caribbean: The Curse of the Black Pearl 66 The Platters: “Smoke Gets in Your Eyes” 224 Poltergeist 99n29 Powell, John 91 Precious 130–132 Promising Young Woman 119–120 Psycho 6, 15, 18, 21n29, 99n27 Punch Drunk Love 26–27, 70–71, 82n12 Purcell, Henry: King Arthur, or the British Worthy 33

FILM, FILMMAKER, COMPOSER INDEX

The Red Violin 146, 151–154, 163, 163n4 Repossessed 134, 135, 142 Return to Oz 89–90 Robinson, J. Peter 56–57 Sakamoto, Ryuichi 102 Saturday Night Fever 158–159 Saving Private Ryan 226–227 Schubert, Franz: “Ave Maria” 129; “Ständchen” 159–161; Unfinished Symphony 120 Scream 99n29 Seal: “Kiss from a Rose” 115n15 Shang-Chi and the Legend of the Ten Rings 114 Sharples, Winston 109 The Shawshank Redemption 53 Shire, David 57, 73, 79, 90 Short Circuit 46 The Simpsons 192n7; “Beware My Cheating Bart” 75; “Them, Robot” 75 Snow White 65 Soundless 44 Spellbound 74–75 Spider-Man 115n15 Stardust 82n8 Star Wars franchise 92, 97n7, 98n16, 107 Star Wars: Episode VI – Return of the Jedi 1 Star Wars: The Return of the Jedi 24–25 Star Wars: The Rise of Skywalker 17–18 The Stepford Wives (2004) 37, 44, 47, 48n11 Stevie Wonder: “Ebony and Ivory” 132 Stokowski, Leopold 164n9 Stothart, Herbert 105 Strauss, Johann: “Blue Danube” 79 Stravinsky, Igor: The Rite of Spring 97n10 Su, Cong 102 Superman (1941) 109 Superman (1978) 111–112, 115n11

247

Super Mario Bros. 115n15 Super Mario video game 115n15 Superstar 129, 131 The Taking of Pelham One Two Three 53–54, 68, 78–79 Tchaikovsky, Piotr I.: “Dance of the Sugar Plum Fairy” 103; “Swan Lake” 121 Ted 136, 137 Terminator 2: Judgment Day 125, 138–140, 143 The Terminator 91 The Three Musketeers (1948) 104–105 Timberg, Sammy 109 Titanic 23–24 Titus 30 Tom and Jerry Greatest Chases: “Yankee Doodle Mouse” 38 The Triplets of Belleville 220, 222n9 The Truman Show 23, 32, 35 Unfaithful 35n7 Up 227–228 Vanessa Carlton: “A Thousand Miles” 129 The Verdict 43 Vertigo 88 Vier Minuten 54–55 Wagner, Richard: Der Ring des Nibelungen 115n17 WALL-E 80, 84n28 White Chicks 129, 130 Who Framed Roger Rabbit? 107 Williams, John 92–94, 102, 111, 113 Wonder Woman franchise 113 Wonder Woman TV series (1970s) 110–111

SUBJECT INDEX

acoustic signatures 167–168, 170n3, 223–224; affective correlates of 167; definition 11; ecological/evolutionary function of 167, 170n4; ethological perspective on 167–168, 170n5; physiological correlates of 167; and timbre 14, 21n22, 21n24 acoustic signatures (film music) 12, 151; and consonance/dissonance (sensory) 14–15; and empathy 10; and subvocalization 11–15, 21n22; and timbre 12–15, 21n24, 148 adaptive cognition 82n9, 167, 200; and leitmotifs 96; and memory 191, 222n7 affordances 63n8, 81, 181–185; and bimodal neurons 183, 186n8; and canonical neurons 183, 186nn7–8; and Cartesian dualism 185n4, 186n5; and cognitive offloading 186n10; ecological/evolutionary perspectives on 81n1, 181; empirical exploration of 182–183; and the environment 81n1, 184; extended 184; and exteroception 185n4; and human behavior 182; micro-affordances 183; and mirror neurons 186n7; of musical instruments 184–185; neural correlates of 183, 186n5; and proprioception 185n4; role in categorization 220, 222n8; of surfaces 182; theory of 181 affordances (film music) 63n8, 64–84, 92, 149–150, 184–185, 228; in the absence of meter 69, 71, 76–77; of alternating meters 68; of asymmetric meters 68, 82n8; of chromatic collection 78–79; cultural entailments of 69, 71, 86; of dodecaphonic pitch collection 73–75; of duple meter 17–18, 32–33, 65–66,

97n7; of duple vs. triple meters 66, 69, 71, 97n7; and entrainment 124n5; and leitmotif 85, 87–89, 97n7; metrical 55, 65–71, 88–89, 124n5, 149; metrical and tonal combined 76–80; of octatonic pitch collection 74–76; of symmetrical pitch collections 73–75, 78, 83n19; taxonomy of 186n11; tonal 72–76, 88, 147–148, 150; of triple meter 66–67, 70–71, 77–80, 82n12, 97n7, 149; of whole-tone pitch collection 72–77 analogues see cross-modal correspondences archetypes 3, 100–115, 150, 203–210, 228; in film music (see topics (film music)); literary 203; in music (see topics (music)) associations (film music) see conceptual blending (film music) attention 96, 171n13, 198; and cognitive load 201n11, 202n13; cross-modal 98n23; and empathy 171n13; to film music 1, 146, 179n3; to music 19n10; rhythmic 81n6; and saliency 191, 198–199, 201n7 Auditory Stream Analysis 194–199, 201nn1–2; and bottom-up/top-down processes 199; and cocktail party effect 198; and innate vs. learned schemas 199 cadences 29, 175, 223–226; Aeolian 111, 115n15; half 124n6; metrical 56–57; and musical syntax 63n11; and SOURCEPATH-GOAL and CONTAINER 56–60, 124n6; and style 211n12; tonal 56–62 categorization 217–221; and affordances 220, 222n8; and Aristotle’s essentialism 221–222n4; attributes in 219–220;

SUBJECT INDEX 249

basic level 220–221, 222n11; and CONTAINER 217; ecological/ evolutionary perspectives on 219, 221n1, 222n5, 222nn7–8; exemplar model of 218–219, 221n2, 222n7; and family resemblance 218; horizontal dimension of 221; neural correlates of 221n2, 222n10; probabilistic approach to 218–219; prototype model of 218–219, 222nn6–7; rule-based approach to 218; and topics (music) 217; values in 219–220; vertical dimension of 220–221 categorization (film music) 125–146; and affordances 144n2; attributes in 125; basic level 143–144, 146n13; and family resemblance 144n1; horizontal vs. vertical dimensions of 144n6; probabilistic approach to 125; prototype model of 142, 145n12; and saliency 125–126; weighted values in 125–126 classical conditioning 3, 85, 95–96, 99n26, 192n13; and memory 188–190 cognitive dissonance: in film music 72–73; and gravity 72, 83n20 conceptual blending 212–215; and the “AHA!” phenomenon 216n5; completion 214; composition 214; and conceptual metaphor 214–215, 216n10; disanalogy 214; elaboration 214; and formal logic 215n1; frames 213–214; and image schemas 214, 216n10; integration 214; mental spaces 212–213; and multidimensional scaling 215n2; music scholarship in 215; neural correlates of 214, 216nn7–8; single-scope networks 216n10; vital relations 214, 216n9 conceptual blending (film music) 116–124, 154–156, 158, 159–163; and concert pieces 120–124; disanalogy 124n2, 160, 162; and formal design (music) 122; and incongruity 124n8; and irony 124n8; and popular songs 116–120; transformations of 159–163; vital relations 124n1, 124n8, 160, 162 Conceptual Integration see conceptual blending conceptual metaphor 50, 154, 172–182; [A] IS [B] structure 47n1, 173; and cinematography 62n1; and conceptual blending 180n14; and the cultural sieve model 175; cultural variation in 175, 180n12; directionality of 47n1, 174, 179n7; and the mimetic hypothesis 180n14; and musical forces 180n14; neural correlates of 179n9, 180n11; origin 174–175; source and target domains of 173–174; theory 173–175; universality of 174–175

conceptual metaphor (film music) 37–63; and CONTAINER 52–62; cultural variability of 48n3; and LINEARITY 37–47; and SOURCE-PATH-GOAL 50–52, 56–62; and troping 154; see also cross-modal correspondences (film music) consonance/dissonance 224–225; and acoustic signatures 10, 14; and behavior 20n16, 21n27; cognitive (see cognitive dissonance); in harmonic constructs vs. in complex tones 15, 21n29, 138; and LINEARITY 43–45, 48n8, 84n27; and masking 21n28; musical vs. sensory 21n23; neural correlates of 21n28; physiological correlates of 21nn27– 28; sensory 83n16, 83n21 contagion (behavioral) 166, 170, 171n14, 171nn16–17; and deindividuation 171n14; emotional 21n30, 22n32 contagion (film music) 6–7, 16–18, 45–46, 149, 228; ecological/evolutionary perspectives on 16; and a human MNS 17 CONTAINER 172–178, 180n17, 217; and SOURCE-PATH-GOAL 50, 177–178 CONTAINER (film music) 23–36, 62nn2–3; and narrative syntax 52; and non-normative sound design 25–35; and normative sound design 23–25; and SOURCEPATH-GOAL 56–62; synchronic vs. diachronic perception of 62n3 cross-modal correspondences 202n13, 228; audio-visual 179n3; and the Congruence Association Model 179n3; cultural variability in 179n5; and developmental psychology 180n10; and the McGurk effect 179n3; and the unity assumption 179n3; vs. intra-modal correspondences 179n4 cross-modal correspondences (film music) 18n2, 37–38, 42, 92, 99n28; and attention 98n23; concomitant relationships 44, 48n12; and consonance/dissonance 43–44; ecological/evolutionary perspectives on 47n3, 48n6, 48n9, 48n12; and loudness 39, 43–45; and pitch frequency 37–41, 43–44, 83n24; semantic 43–47; and sound’s envelope 51–52; and static vs. dynamic parameters 40–41; structural 37–42; tempo and 41–42; and timbral density 44–45; and timbre 46–47 diegesis 10, 23–34, 35nn1–7, 36n9, 36n16, 41, 46, 48n11, 52, 62, 63n6, 63n14, 67, 69–71, 116–117, 134, 140, 150, 158, 226–289; and CONTAINER 23, 35n2; definition of 35n1 ecological resonance see perception (auditory) embedded mind 185n2

250

SUBJECT INDEX

empathy (film music) 6–22, 45, 86, 228; and consonance/dissonance 20n16; and contagion 16–18, 149, 45–46 (see also contagion (behavioral)); ecological/ evolutionary perspectives on 22n32; entrainment and 8–11 (see also entrainment); and a human MNS 7, 18, 19nn5–6; neural correlates of 19n5; physiological correlates of 19n7; process model of 6–7, 18, 18n1; and selfregulation 19n11; and subvocalization 11–15, 45–46 (see also subvocalization) empathy (social) 82n11, 165–170; and electromyography (EMG) 171n13; evolutionary function of 170; and gestures (physical) 166–167; and gestures (vocal) 167; and a human MNS 169, 170n2; and inhibitory mechanisms 169–170, 171n13; and mimesis 166; neural correlates of 166, 169, 170nn1–2, 171n12; and postures 165–170; process model of 18n1, 165–170 entrainment: ecological/evolutionary perspectives on 82nn9–11; neural correlates of 8, 19nn8–9, 81n6; physiological correlates of 20n13, 82n13; and saliency 21n30; and speed thresholds 81n2 entrainment (film music) 6–11, 18, 45–46, 49n15, 81n2, 92, 124n5, 228; and affordances 65; and associations 69; and cardiorespiratory rates 20n12; ecological/ evolutionary perspectives on 82n10; of meter 66–71; for physical labor 81n4; physiological correlates of 20nn12–13, 81n6, 82n7 envelope (sound) 62n1 ESMAMAPA 2, 3, 4, 223, 228 formal design (film music) 52, 227–228; binary 52–55; one-part 52–53; rondo 52–54; sonata 122, 124n7; variation 62 gestalt principles 182; common fate 197; figure-ground 198; good continuation 197; masking 197; and memory 188; proximity 196; and saliency 94, 198; similarity 196 gestalt principles (film music) 97n9; common fate 93; proximity 92–93; similarity 92–93 gestalt principles (music): common fate 197; figure–ground 198; good continuation 198; masking 198, 201n6; proximity 196; and saliency 198; similarity 196–197 gestures (film music) 6–7; cultural origin of 19n3; physiological correlates of 6–7; and topics (music) 19n3

gravity: and cognitive dissonance 72; in film 72–73, 83n20, 84n26; in music 72–73; physical 72; and proprioception 72–73 image schemas 172–182; BLOCKAGE 63n10; CONTAINER 175–176; LANDSCAPE 175; LINEARITY 175–177, 180n17; neural correlates of 180n11, 180n17; SOURCE-PATH-GOAL 175, 177–178, 180n17; VERTICALITY 175–177, 180n17 instrumentation see timbre irony see rhetorical devices leitmotif 64, 226; and affordances (film music) 85, 87–89, 97n7; as classical conditioning 85, 95–96, 99n26; construction of 85–92; and cue abstraction 97n9, 98n18; definition 85; ecological/evolutionary perspectives on 86, 94, 96, 99n29, 163n1; and embodied schemas 147–150; in film musicology 96n1; and gestalt principles 92–94, 97n14; and imprint formation 97n9, 98n18; and memory 85, 94–95; neural correlates of 97n14; and onomatopoeia 85, 89–91; in opera 98n23; perception of 85; and saliency 94, 97nn12–14, 99nn27–28; and topics (film music) 91–92, 147, 150–154; transformation of 146–154 LINEARITY (film music) 37–49, 83n24; see also cross-modal correspondences memory 187–192; and amnesia 191n4; and bottom-up vs. top-down processing 188–189; classical conditioning 189, 192nn12–13; connectionist model of 191n5; echoic 187–188; ecological/ evolutionary perspectives on 190–191; episodic 190; and expectation 193n16; and gestalt principles 188; implicit vs. explicit 189, 192n11; and language 99n24; long-term 189–190, 192nn6–7; of melodies 98n14; modal vs. amodal representations of 192n15; neural substrates of 192n7; phonological loop 188; priming 189; procedural 190; record-keeping theory 187–188; and saliency 191; and schemata 190, 192n9; scripts 190; semantic 190; sensory 187–188; short- vs. long-term 188; and subvocalization 188; working 188–189, 192nn6–7 memory (film music) 85–99, 98n15, 228; bottom-up vs. top-down processing 94–95, 97n9, 97n14; classical conditioning 95; and cross-modal attention 98n23; cultural variability of

SUBJECT INDEX 251

94; implicit vs. explicit 95; and leitmotifs 95, 98n16; phonological loop 95; and tonality 98n20 metaphor 173–175; and categorization 178n1; and cross-modal correspondences 172–173, 178n2, 179n3; see also conceptual metaphor Mickey Mousing 38, 42, 67 mirror neuron system (MNS) 20n20, 165, 169–170; audiovisual mirror neurons 169, 171n11; and behavioral contagion 170; as bimodal neurons 186n8; vs. canonical neurons 186n7; and emotion 171n12; and empathy (social) 169; and film music 7, 18; in humans 170nn9–10; music-based 7 modes see pitch collections onomatopoeia 27, 30, 34, 36n14, 65, 148–149; definition 89; semiotic perspective on 87, 97n3 orchestration see timbre perception (auditory) 194–201, 228; and Auditory Scene Analysis 194–195 (see also Auditory Stream Analysis); bottom-up vs. top-down processing 194–195, 201n2; ecological/evolutionary perspectives on 194, 199–200; and gestalt principles 194–195; and innate schemas 194–195, 199, 201n3; and learned schemas 194–195, 199, 201n3 perception (film music) 85–99 perception (visual) 196–199; bottom-up vs. top-down processing 200; ecological/ evolutionary perspectives on 200 pitch collections 91, 152; Aeolian 92, 103–104, 106–107, 112, 114, 119, 150–151, 207–208; chromatic collection 78–79; dodecaphonic 73–75; Dorian 62n4, 151; Hungarian minor 103; Ionian 109, 115n10, 161, 204; and leitmotifs 91; Lydian 80; octatonic 74–76; pentatonic 91, 97n5; Phrygian 60, 62, 100; qualia of 83n18, 83n25; symmetrical 73–75, 78, 83n19; and tonal centricity 83nn18–19; whole-tone 72–77 proprioception 72–75, 83n16, 83n20, 170n7; and affordances 72, 185n4 qualia: of harmonic progressions 115n17; of pitch collections 83n18, 83n25; of scale degrees 113, 115n18 red herring 48n10 rhetorical devices 228; categorization of 126–127, 144n2, 146n13; irony 124n8,

125, 127; lie 126–127; paradox 125, 127; parody 125, 127, 144n6, 144n8; quotation 126–127; sarcasm 126–127; satire 125, 127, 144n6, 144n8 rhetorical devices (film music): irony 128–136, 140–141, 144nn9–11; lie 138–141; paradox 138, 142; parody 135–139, 141–142, 144n11; quotation 137, 142; sarcasm 121, 133; satire 131–134, 138, 141, 144n11 saliency: in conceptual metaphors 179n7; neural correlates of 201n8; and schemas 201n10 saliency (film music) 97nn12–14, 98n18; and categorization 125; and the Congruence Association Model 202n13; and memorization 99n28 schema see image schema semiotics: Anglo-American 206–207; chain of significations 207; continental 206–207; denotation/connotation 206–207; icon 207; index 207; Peircean 206–207; signifier/signified 206–207; structuralist 206–207; symbol 207 semiotics (music) 206–209; icons vs. indexes 114n5; and leitmotifs 88–89; and topics (music) 203, 207–209 Shepard tone 74, 83n23 signs see semiotics sound design 62n5; and CONTAINER 23, 62n3, 63n17 (see also CONTAINER (music)); and cultural variability 36n9; non-normative 25–35; normative 23, 35n3; overlap 25–26; replacement 27–30, 63n14; and technology 36n8; transference 10, 30–34, 48n11, 63n6, 63n17, 124n6, 134, 150 SOURCE-PATH-GOAL see image schemas SOURCE-PATH-GOAL (film music) 50–63; and CONTAINER 56–62; and narrative syntax 50–52 subvocalization 92, 95; and memory 188 subvocalization (film music) 6–7, 11–15, 16, 45–46, 151; and a human MNS 12, 17; neural correlates of 11, 20n19, 20n20; physiological correlates of 11, 20nn14– 15, 20n18, 21n22 timbre 63n9, 87; associations of 63n7, 113; and distortion 46; and onomatopoeia 148 (see also onomatopoeia); and saliency 94 tonal centricity: and pitch collections 72–73; and rare intervals 83n19 topics (film music) 92, 97n7, 100–115, 137–140, 147, 151–159, 225; diachronic perspective 108–113; function 100; and leitmotifs 91–92, 114n1; and listeners’

252

SUBJECT INDEX

competence 150–151; markers 155; synchronic perspective 100–107; and timbre 101 topics (music) 203–210; in 18th-century 203–204, 211n4; and canned music 211n13; categorization of 217; cultural variability of 210; definition 91, 100, 150–151; empirical perspective on 209, 211n6; and ‘feels’ 204; in film 205, 208, 209, 217; and formal design 211n6; and

gender 210; juxtaposition of (see troping); and listeners’ competence 204–205; markers 208–210, 210n1, 211n5; neural correlates of 209; in popular culture 204, 210; semiotic perspective on 206–209, 210n1; and sentic types 211n9; in silent films 211n11, 211n13; and stereotypes 210; Universe of Topics 203–204; in video games 211n14 troping 154–159; definition 163n5