+ All Categories

fpaper

Date post: 05-Jul-2018
Category:
Upload: juan-perez-arrikitaun
View: 214 times
Download: 0 times
Share this document with a friend

of 27

Transcript
  • 8/15/2019 fpaper

    1/27

    Algorithmic Composition

    Andrew Pascoe

    December 7, 2009

    1 Introduction

    Advances in computing technology have enabled composers to use the computer as

    a means of generating musical content. Some of this musical content is presented

    as original, and some of it is presented as extending particular composers’ styles.

    Regardless of how it is presented, algorithmic processes have caused controversy in

    traditional music circles. How is algorithmic music authentic? How can algorithmic

    music have meaning? Are human composers destined to become obsolete? How do

    these techniques relate to the question of man versus machine?

    This paper begins with a discussion of the long history regarding the systematiza-

    tion of music. These systematic approaches to music theory provide a foundation for

    the possibility of computers to produce “good” music. Only some brief examples are

    provided. As such, understanding the role of systematization in music is necessary to

    have an intelligent dialogue about this one direction music is heading.

    From there, the paper covers some recent methods and approaches to producing

    algorithmic compositions. In particular, the works of Benjamin Carson, David Cope,

    Peter Elsea, and Microsoft’s MySong  application are examined. Each of these attacks

    the problem of automatically generating music in different ways.

    Last, the paper discusses musical meaning and the philosophical questions at-

    tached to this concept. These are broad questions with perhaps no correct answer.

    1

  • 8/15/2019 fpaper

    2/27

    However, they serve to provide at least some notion of why algorithmic composition

    is a controversial subject. The role and intent of the composer form the basis for this

    discussion.

    2 Historical Systematization of Music

    2.1 The Sixteenth Century

    Consider the following piece of music:

     

      2 4

      Figure 1: Bad counterpoint.

    This fragment is considered to be poor compositional form, even today.[9] The

    technical reason is the  hidden octave  on the second beat. While there is no question

    that composers in the sixteenth century functioned under at least the intuition of 

    such rules, the real formalization of these practices was completed by Johann Fux in

    his  Gradus ad Parnassum  in 1725.[15] As Taruskin describes it:

    [Musicians] used it to gain facility in “the first principles of harmony and

    composition,” which were regarded by teachers as an eternal dogma in

    its own right, a bedrock of imperishable lore that “remains unaltered, let

    taste change as it will.”

    The text itself divides counterpoint into five   species , along with set of rules con-

    cerning prohibition against perfect parallels, hidden parallels (as above), consonances,

    dissonances, and also general aesthetics such as restrictions on sequential parallel

    thirds and sixths, and a tendency away from parallel motion. These were derived

    2

  • 8/15/2019 fpaper

    3/27

    from studying the works of Palestrina. Modern composition students still receive

    instruction in these rules and aesthetics.

    Already one can begin to see how a computer can begin to use these notions of good

    composition versus bad composition. It is ultimately simply a matter of eliminatingpossibilities and ranking the remaining possibilities in terms of their adherence to the

    given aesthetic.

    2.2 The Common Practice Period

    The common practice period encompasses most styles associated with the term “clas-

    sical music:” Baroque, Classical, and Romantic period styles. The theoretical ground-

    work for these styles is based not only in the counterpoint and voice-leading consid-

    erations of Fux, but also in chord progressions. Kostka and Payne write:[11]

    [Students] must learn which chord successions are typical of tonal harmony

    and which ones are not. Why is it that some chord successions seem to

    “progress,” to move forward toward a goal, while others tend to wander,

    to leave our expectations unfulfilled?

    They provide the following examples:

     43

      43

     

       

    Figure 2: Tonal harmonic progression.

    3

  • 8/15/2019 fpaper

    4/27

     43

      43

     

       

     

    Figure 3: Random harmonic progression.

    The implication is that there exists a set of guidelines for writing good, tonal music.

    These guidelines are what is typically referred to as “music theory,” in imprecise,

    casual speech. Kostka and Payne summarize some of these guidelines with a diagram

    of possible chord progressions, but are quick to point out:

    . . . [B]e aware that Bach and Beethoven did   not   make use of diagrams

    such as these. They lived and breathed the tonal harmonic style and had

    no need for the information the diagrams contain. Instead, the diagrams

    represent norms of harmonic practice observed by theorists over the years

    in the works of a large number of tonal composers.

    That is to say, these rules and guidelines have been imposed  after the fact, much asFux’s Gradus ad Parnassum  did for the sixteenth century style. Once again, this push

    toward systematization continues to reduce the art of composition to an algorithmic

    framework.

    2.3 The Twentieth Century

    The twentieth century is unique in at least one respect to previous centuries: In ad-

    dition to theorists imposing theoretical structures on previous composers, composers

    themselves were creating systems for generating music. This section provides a couple

    examples from each camp.

    4

  • 8/15/2019 fpaper

    5/27

    2.3.1 Charles Seeger

    Charles Seeger, in a twist of Fux’s work, devised a systematic approach to composition

    he dubbed “dissonant counterpoint.”[14] Nicholls offers the following example:

        

        

     

    Figure 4: Dissonant counterpoint.

    Here, the focus is on dissonance as opposed to consonance. This example has no

    consonances between the two parts. Moreover, the individual lines themselves have

    a tendency toward dissonant intervals, e.g. the B to F in the top line is a tritone.

    Repetition of pitches is to be avoided unless they are sufficiently far apart. As the

    number of voices increased, rhythmic dissonance also becomes a necessity. All in all,

    these rules and guidelines codify a means of creating a new type of music.

    2.3.2 Arnold Schoenberg

    Arnold Schoenberg created a technique that is known as “serial composition,” “twelve-

    tone technique,” or “dodecaphony.”[16] This atonal process begins with a tone row of 

    all twelve diatonic pitches. The motivation for this practice is to have the pitches not

    relate to any particular key, but only to themselves. From here, only certain musical

    transformations are allowed: transposition and inversion.   P   is used to denote the

    “prime” tone row. Transposition is denoted through the use of subscripts, so, for

    example, a transposition of six semitones would be written as P 6

    . Inversion is denotedby   I   , and inversions can also be transposed. Thus, an inversion transposed by six

    semitones would be written as  I 6. Taruskin gives the following examples using a tone

    row from Schoenberg’s  Suite, Op. 25 :

    5

  • 8/15/2019 fpaper

    6/27

                     (a)  P 0

                    

    (b)   I 0

                        (c)  P 6

                             (d)   I 6

    Figure 5: Twelve-tone processes.

    These processes are extremely mathematical in nature. Besides just being a gen-

    eral sense of what is “good” composition versus what is “bad” composition, the rules

    laid forth by Schoenberg comment on what is  possible   composition. Moreover, there

    is no question that Schoenberg intended this system to function as hard and fast

    rules. Taruskin writes, and perhaps provides a snide comment toward algorithms as

    a compositional practice:

    Indeed, the use of an exhaustive twelve-note series makes the  Grundgestalt 

    function, and the organic unity thus guaranteed, virtually automatic. And

    that, of course, was the great breakthrough, the “principle capable of 

    serving as a rule,” which allowed the composition of large-scale, abstract,

    and autonomous atonal music of constant and at-all-time-demonstrable

    motivic coherence despite its renunciation of predefined tonal hierarchies,

    and despite its frequent “athematicism.” But beware! As soon as any

    musical characteristic becomes the automatic result of a method, it stops

    being a compositional achievement.

    Regardless of Taruskin’s personal opinions, Schoenberg’s serialism does provide

    a rigorous framework that a computer can be easily made to emulate. Whether a

    computer can make a “compositional achievement” is a question saved for a later

    section of this paper.

    6

  • 8/15/2019 fpaper

    7/27

    2.3.3 Leonard Bernstein

    Leonard Bernstein began to pose new theoretical question related to the structure

    of music, and in that sense his theory imposes interpretations on compositions from

    the past.[2] His theory has its roots in the linguistics of Noam Chomsky. The Chom-

    skian framework of “transformational linguistics” is applied to music. Essentially,

    Bernstein equates melody with basic musical meaning, chords as adjectival modifiers,

    rhythm as verbs, and mode as negation. He continues this reasoning and explores

    the syntactic structures of music by suggesting an underlying “deep structure” to

    musical content that goes through a sort of transformational grammar through not

    only these modifiers, but also grammatical aspects of deletion and combination. He

    freely admits that his ideas are not fully formed, but here are some basic examples

    with potential linguistic equivalents:

    (a) Jack loves Jill.

     (b) Jack doesn’t love Jill.

        (c) Does Jack love Jill?

     (d) Doesn’t Jack love Jill?

    Figure 6: Transformational musical grammar.

    This notion is very different from the forms of music theory we have thus far

    considered. It does not comment on notes or chords in particular, but rather focuses

    on the construction in form of a piece of music. Bernstein likens this to elevating

    prose into poetry.

    Bernstein’s analysis lends a different type of insight into the process of composi-

    tion. This potentially affords computers the opportunity to replicate human creativity

    on a separate level. By using musical fragments and applying transformational gram-

    7

  • 8/15/2019 fpaper

    8/27

    mars to them, new creative structures may emerge. This question is related to natural

    language processing.

    2.3.4 David Lewin

    David Lewin made advances in music theory with his work in generalized musical

    intervals and transformations.[12] In many ways it is an extension of Schoenberg’s

    atonal practice, but through this extension finds relevance to tonal composers, such

    as Chopin. Lewin’s theorization of music relies heavily on the mathematical field of 

    abstract algebra. Indeed, Lewin’s book reads more like a math text than a treatise on

    music, containing theorems, corollaries, and proofs. In the foreword, Edward Gollin

    sums it up succinctly:

    The work, a methodical examination of the concept of a musical interval,

    explores how the familiar notion of interval as “a distance extended be-

    tween pitches in a Cartesian space” is merely one specific case of a more

    general idea, one that can embrace different kinds of musical objects (du-

    rations, meters, Klangs , timbres, and so on), different (i.e. non-Euclidean)

    geometries, and different orientational perspectives (interval as action or

    gesture rather than as simply measurement of distance between things).

    By far the most mathematical example covered in this paper, Lewin’s work finds

    a fit in the world of computation. Not just composition, but   sound itself   can be

    analyzed rigorously within Lewin’s framework. This opens the door for computers

    to not only compose well, but to also construct interesting collections of frequencies

    that evolve over time for synthetic instrument creation—aspects of music that are

    hardly intuitive for traditional composition practitioners. Thus, the computer can

    more easily express musical ideas than a human can.

    8

  • 8/15/2019 fpaper

    9/27

    3 Recent Methods

    This section of the paper now looks at approaches to algorithmic composition from a

    few perspectives, namely Benjamin Carson, David Cope, Peter Elsea, and Microsoft’s

    MySong  application.

    3.1 Peter Elsea

    Peter Elsea has explored algorithmic composition through the use of fuzzy logic

    systems.[8, 7] Pitch classes are represented as sets that denote membership in a par-

    ticular chord or scale. For example:

           (1 0 1 0 1 1 0 1 0 1 0 1)

    Figure 7: Fuzzy logic set of notes.

    This example in the major scale does not contain any real fuzzy logic (though the

    set still can function as a fuzzy logic set). The example of the minor scale, in which

    the seventh can be raised, provides a fuzzier set:

                (1 0 1 1 0 1 0 1 1 0 0.7 0.6)

    Figure 8: Fuzzier logic set of notes.

    Note that these are not probabilities, but merely demonstrate a preference for the

    flatted seventh over the naturalized one.

    This fuzzy logic system affords a host of mathematical operations that can be per-

    formed. Transposition is accomplished by circularly rotating the set. Multiplication

    9

  • 8/15/2019 fpaper

    10/27

    between sets is available. Set unions and intersections are accomplished by taking the

    maximum and minimum, respectively, of two numbers with the same index. These

    operations allow for simple processing in tonal harmony applications.

    Elsea’s software functions by reading in a MIDI note and outputting a harmoniza-tion for that note. The basic set of rules are, with regards to chord inversions:

    If root position keeps common tones, then root position.

    If first inversion keeps common tones, then first inversion.

    If second inversion keeps common tones, then second inversion.

    If last position was root, then first inversion or second inversion.

    If there have been too many firsts in a row, then root or second.

    If there have been too many seconds in a row, then root or first.

    If last position was not root, then root.

    The first three rules ensure that the harmonization does not fly around too wildly.

    The next three rules ensure that the harmonization is not too static. Of particular

    note for these rules is the use of the fuzzy concept of “too many.” Thus, programmed

    into Elsea’s code is a sense of what “too many” actually means. The last rule has a

    relatively low weight of being implemented so that it does not interfere with the main

    rules too much.

    The question, once we move beyond simple inversions, is how to harmonize melodic

    notes given their context not only in the key, but also in the context of the preceding

    chord. Elsea constructs three candidate chords by using these aforementioned fuzzy

    logic sets: one with the note as the root, one with the note as the third, and one

    with the note as the fifth in the chord. Then, the actual roots of these chords are

    compared with more fuzzy logic sets that are based on the tonal harmony (the dia-

    grams) provided by Kostka and Payne. The best candidate, based merely on taking

    the maximum of the resultant computations, is then constructed and played. So for

    10

  • 8/15/2019 fpaper

    11/27

    example, in the key of C, if the user plays a G, there are three basic triadic chords in

    the key of C that harmonize G, not including their inversions:

     V  

    I     iiiFigure 9: Triadic chords in C that harmonize G.

    Which chord to choose is, once again, dependent on the context. If the preceding

    chord is a root position I chord, then the following is a possibility:

    I(6)

    I

    Figure 10: One possible progression.

    But if the preceding chord is a ii chord, the next two examples show a “good”

    progression and a “bad” progression:

     V(6,4)

    ii

    (a) Good progression.

    iii

    ii

    (b) Bad progression.

    Figure 11: Progressions of varying favoritism.

    In short, Elsea takes a mathematical approach to the tonal theory of the common

    practice period, and thus the computer is able to make informed, real-time compo-

    sitional decisions based on melodic user input. This would not have been possible

    without a systematization of the music of the period in the first place.

    11

  • 8/15/2019 fpaper

    12/27

    3.2 Benjamin Carson

    Benjamin Carson, in his compositional practice, takes an algorithmic approach to

    rhythm.[4] The software he has developed in collaboration with a software engineer

    focuses on syncopation over an underlying music pulse, as opposed to the practice of 

    new complexity featuring such methods as nested tuplets. The software is designed

    to give the user a fine-tuned control of multiple aspects of the composition process.

    Fundamentally, the software develops an aesthetic based on the complexity of 

    durations of musical events. For example, a series of musical events may have metric

    times of    416

    ,   716

    , and   1316

    . These durations are randomly generated, and depend upon

    the composers’ input as to how complex he wishes a certain series of rhythmic events

    to be.

    Beyond this, the software allows a composer to view his work from a variety of dif-

    ferent dimensions. Rhythmic complexity can be tied to pitch classes or dynamics, for

    example. Thus, even though the underlying algorithmic production of these durations

    is inherently random, the composer is free to choose some sort of logical framework for

    how rhythmic complexity interacts with the rest of the composition’s aspects. This

    affords a unique mix of both the computer’s input and the compositional goals and

    ideas of the composer.

    These syncopated rhythmic grammars can be variably juxtaposed to the underly-

    ing ordered pulse of the music. This can be accomplished by shifting the syncopation

    over the underlying pulse, or by spreading out the underlying pulse in clock time.

    These processes generate psychological musical effects for the listener. Once again, it

    is at the composer’s whim when to produce such rhythmic effects as part of the cre-

    ative process. In particular, Carson describes this feature of his software as a means of 

    maintaining interesting rhythmic flows that do not grow stale, but still maintain some

    sort of logical structure. The goal is not to be completely random, but instead to have

    a certain level of ordered randomness. Thus, the piece has a concrete development

    over its course without devolving into either banal repetition or pure chaos.

    12

  • 8/15/2019 fpaper

    13/27

    Carson also makes a distinction between metric time and clock time. That is, a

    32nd note does not lose its “32nd-ness” under tempo changes. Its musical function

    does not change despite its variation in actual clock time. To this extent, Carson also

    experiments with an algorithmic approach to changes in tempi.Carson’s software is in at least part a practice in “compositional economy.” By

    automating certain musical processes, he finds himself free to focus on other aspects

    of his composition he finds either more interesting or more important. However, the

    software allows for a large range of input by the composer, allowing him to express

    his “musical impulses” without necessarily having the computer completely control

    the output. Carson’s idea of musical impulses will be covered in the next section of 

    this paper.

    3.3 David Cope

    David Cope has long been involved in the field of algorithmic composition, and much

    of his focus has been on making computer models of musical creativity.[6] His pro-

    gram, Experiments in Musical Intelligence (EMI), employs a variety of techniques to

    accomplish not only algorithmic composition in the styles of particular composers,

    but also to produce new, “creative” output.

    Cope’s approach is fundamentally database driven. He compiles musical sources

    from a particular composer (or particular composers), all of which go through a series

    of analyses. From here, the program sets to work using models of recombinance,

    allusion, learning, form and structure, and influence. This paper presents a summary

    of this ideas and their function.

    3.3.1 Recombinance

    Recombinance is essentially taking musical material from the database and replacing

    sections of it with other material acquired from the database. Cope’s book provides

    a simple example of this process:

    13

  • 8/15/2019 fpaper

    14/27

      43 

    8

     43 

     43 

     43

     

    (a) Original Bach.

     43 

       43

       43

    8

       43

     

    (b) EMI output.

    Figure 12: Recombinance in action.

    The EMI output replicates the original Bach Chorale no. 188 on the first beat,

    but diverges from there, instead replacing it with a fragment from Bach’s Chorale

    no. 157. EMI accomplishes this by forming examining the database and searching for

    alternative candidates following from the given first beat. This process is continued

    throughout the algorithmically generated composition.

    However, this process will ultimately result in free-flowing melodies without any

    sense of direction. Cope writes:

    In order to provide some sense of this logic and larger structure, and again

    wanting to avoid coding my own knowledge of musical form as rules, I

    rewrote the program to inherit more of the structural aspects of the music

    in its database. This inheritance involved extending the analysis process

    so that the program could store other information about the music being

    analyzed in each grouping along with destination notes. For example, I

    had the program analyze the original music’s distance to cadence, position

    of groupings in relation to meter, and other context-sensitive features.

    The program is also sensitive to “compositional signatures,” which are larger frag-

    ments of music that give a particular composer a sense of musical identity. Thus, there

    14

  • 8/15/2019 fpaper

    15/27

    is an aspect of the program that resists the complete fragmentation of raw recombi-

    nance. Due to similar concerns of over-fragmentation, EMI also tries to maintain a

    cohesion in musical texture by comparing density of notes and rhythmic characteris-

    tics between groupings, and searching the database for passages of similar quality intexture.

    3.3.2 Allusion

    Allusion refers to the practice of composers to borrow material from other composers.

    Cope breaks down allusion into five categories: quotations, paraphrases, likenesses,

    frameworks, and commonalities. He freely admits that the boundaries between these

    categories are not necessarily clear.

    Quotations are as one would expect: direct pitch and rhythm copying. Para-

    phrases use similar intervallic structures to the source material and can also incor-

    porate rhythmic differences. Likenesses are similar to paraphrases, but allow more

    freedom in the differences, as well as changes in harmonic content. Frameworks take

    source material and insert notes between the pitches and rhythms of the source mate-

    rial. Such processes can be illuminated through textural reductions. Commonalities

    are simplistic musical materials, not unlike Bernstein’s concept of “deep structures.”

    Cope developed a secondary program called Sorcerer that can sniff out these allu-

    sions. A user provides Sorcerer with a database of musical material and a work that

    potentially contains allusions that are in the database. The program compares not

    only pitches but intervals between pitches, allowing for variation in these parameters,

    and also allows for a certain number of notes to fall within the source material (the

    framework idea). Users can specify how sensitive they wish Sorcerer to perform, pick-

    ing out only quotations, or running the whole gamut up to commonalities. Bounds

    on the length of patterns must also be set by the user.

    EMI, because of its construction in accessing databases, is naturally inclined to at

    least include some elements of allusion. Cope tells a story about how he recognized

    15

  • 8/15/2019 fpaper

    16/27

    in one of EMI’s pieces an allusion to Vivaldi, despite the database being entirely

    comprised of works by Bach. Bach himself transcribed a series of Vivaldi pieces. Did

    EMI reveal a hidden allusion in Bach to the music of which he himself had intimate

    knowledge?

    3.3.3 Learning

    Cope describes another program he worked on called Gradus (named after Fux’s

    Gradus ad Parnassum ) which attempts to learn the rules of first species counterpoint.

    Other programs in the past have codified solutions to this problem, but they do not

    learn in particular.

    A melodic line is given to these programs (called a  cantus firmus ). The goal is to

    write suitable counterpoint against this line. Previous programs begin with the first

    note and proceed to the next. As any student of counterpoint can attest to, this naı̈ve

    approach can quickly run into serious problems. Some beginning solutions resolve in

    dead ends where there are no possible notes to write that will conform to the rules.

    Therefore, such a program needs to backtrack to the last good note and attempt to

    replace it with another possible solution, and then continue on from there, potentially

    running into the same problem all over again.

    Cope’s program stores its mistakes and avoids replicating them again in the future.

    So the program starts by having to backtrack, but the more the program “practices,”

    the less it must backtrack to find a solution to the given cantus firmus. While Cope’s

    example has limited application to sixteenth century first species counterpoint, the

    underlying philosophy has large implications for algorithmic composition as a whole.

    The ability of programs to learn what to avoid can not only increase efficiency, but

    also produce higher quality work.

    16

  • 8/15/2019 fpaper

    17/27

    3.3.4 Form and Structure

    Cope considers structure in the form of SPEAC analysis: statement, preparation,

    extension, antecedent, and consequent. These all refer to relatively brief musical

    gestures that find their characteristics in the amount of tension they provide to the

    listener’s ear. The interactions between these different aspects of music create logical

    surface structures. Alas, Cope’s book does not go into detail about how such an

    analysis is implemented in his programs. He writes, “I simply ask readers to imagine

    the possibilities, and to investigate the realities on their own.”

    As for form, SPEAC analysis is applicable. Cope outlines another program called

    Alice that analyzes a database searching for broad compositional forms. It begins

    by attempting to identify themes. If this attempt is unsuccessful, it analyzes texture

    changes to deduce the form. Failing that, it relies on simple cadences. If that process

    also fails, Cope says Alice “takes its best guess.” From what little details Cope gives

    in his book, it appears Alice searches for “expectation, fulfillment, and deception of 

    themes or progressions,” and tries to replicate these structures in its database.

    3.3.5 Influence

    Influence is separate from relying on a database or merely composing in another’s

    style. Instead, influence is about creating a new style that has its roots in a previous

    style. Cope gives three examples of his experiments with EMI and musical influence.

    The first example is a feedback loop. EMI begins with a database of a particular

    composer’s work. As has already been described, it produces compositions based

    on this database. Once a composition is completed, it is inserted back into the

    database. By repeating this process thousands of times, Cope found that the output

    showed obvious “influence” from the original source material, but the overall style

    was exceptionally different.

    The second example uses multiple EMI programs all using databases from different

    composers. Here, the instances of EMI communicate with one another about their

    17

  • 8/15/2019 fpaper

    18/27

    compositions. Cope programmed in a metric for what a “good” composition is: the

    kinetic energy of the music. Each program has a slightly different metric, and hands

    over to the other programs music it evaluates as aesthetically acceptable. In this way,

    the programs are attempting to influence each other, resulting in compositions thatdemonstrate disparate sources. The author finds a similarity between this idea and

    Bohm Dialogues.

    Last, Cope describes influence as a form of exploration. He created a web-crawling

    application called Serendipity that would download MIDI files (and text files) that

    would be added to databases. At first, this process resulted in a mish-mash of styles

    in one composition that was not aesthetically pleasing. So Cope produced filters

    that would determine if a given MIDI file was good enough, or even appropriate,

    for inclusion in EMI’s database. This resulted in much more cohesive and enjoyable

    compositions.

    3.4 MySong

    MySong   is a consumer application that has been produced by a Microsoft research

    team.[13] While its application is not as serious as what has been previously exam-

    ined, the author felt its inclusion to be beneficial as an example of giving a “taste

    of composition” to musically untrained individuals. That is to say, algorithmic com-

    position is becoming a tool for users to create their own content. It also functions

    differently than other work presented in this paper.

    However, there are some similarities to the other work. Like Elsea’s work,  MySong 

    attempts to harmonize melodic input by users, and like Cope’s work, it does so

    by using a database of music. The distinctions lie in the implementation of these

    practices.

    To begin with, users specify a tempo and are provided with a metronome track

    that delimits bar lines, or measure boundaries, in their melodic input. Users then sing

    a melodic line into a microphone. The software then uses a pitch detection algorithm,

    18

  • 8/15/2019 fpaper

    19/27

    and takes all of the notes in a particular measure and compiles them into a set. Here

    is an illustration:

        (a) Sung line.     (b) Pitch classes.Figure 13: Pitch sets derived from melodic lines.

    The database MySong  employs is a collection of about three hundred lead sheets.

    Lead sheets contain melodic content along with chord progressions. These lead sheets

    come from a variety of popular music genres, including pop, rock, and country. As

    such, these styles of music usually find changes in harmony at the bar line. Here is

    an example of what a lead sheet looks like:

    G

      C

      F

    Figure 14: Simple lead sheet example.

    The method for algorithmic composition is based in Hidden Markov Models (HMM).

    Essentially, a Markov model represents a stochastic process through transition prob-

    abilities. Based on an analysis of the database, we might conclude that a C major

    chord will move to a G major chord 30% of the time, and move to an F major chord

    20% of the time. What makes the Markov model hidden in this application is that

    the chord progression is unknown—all that is available are the pitch classes derived

    from the user input. Thus, there is another level of analysis involving the probability

    of a chord for the measure given the pitch class.

    The way to find the optimal solution given a set of observances in an HMM is to

    use the Viterbi algorithm. A mathematical discussion of the Viterbi algorithm and

    why it works is beyond the scope of this paper, but it is important to know that there

    19

  • 8/15/2019 fpaper

    20/27

    is an input parameter  k  to which the users of  MySong  are given access.

    This   k  is exhibited in  MySong  as the “Jazz Factor.” If   k  = 0, then the program

    effectively ignores changes in the melodic line altogether and produces the most typical

    chord progression deriving from the first measure. If  k = 1, then the program ignoresstandard chord progressions and the harmonization for each measure is determined

    solely by the pitch class for that measure. That is, when the Jazz Factor is increased,

    the program is more likely to make bold harmonic choices. If  k  = 0.5, then for each

    measure the program equally accounts for the user’s input and the chords in the

    preceding and succeeding bars.

    The other main user control in   MySong   is the “Happiness Factor,” which also

    ranges from 0 to 1. Chord progressions for major and minor modes differ, and the

    application accounts for this. When   h   = 0, the chord progression is guaranteed to

    be in the minor mode, and when  h = 1, the chord progression is guaranteed to be in

    the major mode. As can probably be deduced, when  h = 0.5, the program will freely

    choose the best chord based on the Viterbi algorithm and the value of  k, whether or

    not it is a major or minor chord.

    Once the Viterbi algorithm is complete, users are then free to tweak their chord

    progressions if they so choose. From there, the user can select a particular style of 

    accompaniment that adds instrumentation and rhythm to the final chord progression,

    and output their song to a sound file. Morris did not elaborate on how this stylization

    is accomplished.

    MySong   is one foray into the public sphere for algorithmic composition. Morris

    notes that given the input of a melody from a famous pop song, the program will at

    least sometimes produce the same chord progression in the studio version of the song.

    He also touts the ability of   MySong   to produce interesting variations on common

    songs by tweaking the Jazz and Happiness Factor controls.

    20

  • 8/15/2019 fpaper

    21/27

    4 Musical Meaning

    Let us consider the following musical fragments:

      42      ff        Figure 15: Opening theme from Beethoven’s Symphony no. 5.

                   

      42               p       

                   

    Figure 16: Subject from Shostakovich’s Fugue no. 5, Op. 87.

        

        43        

     Figure 17: Opening output of EMI of a Bach-style lute suite.[6]

    What, if anything, do these bits of music mean? Taking a cue from  MySong ’s

    use of a Happiness Factor, should we conclude that Beethoven was having a bad day,

    and Shostakovich happened to be feeling jubilant when they wrote their respective

    pieces? One of the author’s past professors suggested that such interpretations were

    dangerous, but Aldwell and Schachter suggest that our connotations of the major

    mode being “happy” and the minor mode being “sad” should not be dismissed too

    carelessly.[1] Of course, this begs the question, “What was EMI feeling?”

    Hofstadter finds these questions, given the quality of EMI’s output, disturbing.[10]

    He writes:

    . . . I was going to grapple with this strange program that was threatening

    to upset the apple cart that held many of my oldest and most deeply

    21

  • 8/15/2019 fpaper

    22/27

    cherished beliefs about the sacredness of music, about music being the

    ultimate inner sanctum of the human spirit, the last thing that would

    tumble in AI’s headlong rush towards thought, insight, and creativity.

    Having a personal affinity for Chopin and being surprised at how well EMI emu-

    lated Chopin, Hofstadter boils his worries down to three possible outcomes, each of 

    which he finds unsatisfactory as a lover of music and as a human being:

    (1) Chopin (for example) is a lot shallower than I had ever thought.

    (2) Music is a lot shallower than I had ever thought.

    (3) The human soul/mind is a lot shallower than I had ever thought.

    These concerns are strange since Hofstadter claims he has long been convinced

    that humans are effectively machines anyway. He claims his surprise lies in the fact

    that the machines producing EMI’s output are many orders of magnitude simpler than

    human biological construction. Regardless, the author is not particularly surprised

    by such developments in the production of machine music—the long historical trend

    of the systematization of music lends itself to mechanical replication.

    But Hofstadter is not alone. Cope has found resistance in music circles to havingEMI’s output performed by trained human musicians, saying responses were mostly

    negative.[5] Why do people have such revulsion to the idea that a computer can

    compose well? Do people read that much meaning into music?

    Bernstein rejects the notion of any sort of extra meaning in music.[3, 2] For him,

    the meaning of music is contained entirely within the notes themselves. He declares

    that music does not have the same function as language, and any meaning it has is

    ineffable. Hofstadter’s claim that music functions as “direct soul-to-soul messages”

    does not find a place in Bernstein’s philosophy of music. As Bernstein says:

    . . . [L]anguage has a communicative function and an aesthetic function.

    Music has an aesthetic function only. For that reason, musical surface

    structure is not equitable with linguistic surface structure.

    22

  • 8/15/2019 fpaper

    23/27

    Cope also rejects Hofstadter’s objections.[6] He writes:

    Interestingly, one of my favorite analogies is that mathematics is to physics

    as music is to language. The first instances—mathematics and music—

    are abstract, while the second instances—physics and language—relate

    more to the real world. Mathematics and music also deal in proportions,

    while physics and language attempt to develop meaning. . . Though sim-

    plistic, this analogy nonetheless serves to emphasize music’s reliance on

    relationships rather than on meaning and representations of meaning.

    Carson, too, suggests that music has no inherent meaning beyond its notes, and

    says he agrees with Bernstein.[4] He states that music is nothing more than a surface

    structure, and it has no need for semiotics. Beyond this, Carson is glad that the study

    of algorithmic composition is finally bringing these issues to light, that the concept

    of “depth” in music is completely misguided, in that it necessarily implies a mystical

    and mysterious element to the composition process that is just simply not present.

    Carson claims that the use of algorithmic processes for compositional practices

    does indeed distance the composer from his work. He also states that the highestlevel of intimacy a composer can have with his composition is when the composer

    specifies each element in the work himself. However, this is not necessarily beneficial.

    For one, as already mentioned previously in this paper, is the issue of economy.

    Carson says that with modern technology, a composer would have to be “crazy” not

    to automate parts of a composition that have a fundamental structure that could be

    mechanically composed. This enables the composer to focus on other aspects of the

    work. Second, a là John Cage, removing the ego of the composer grants opportunities

    for musical exploration to which the composer’s mind may never have led him.

    In response to a follow-up question relating to economy, that perhaps the most

    economical action to take is to simply rely on algorithmic composition for all parts

    of the process, Carson says that he sees no real issue in this. Composition, he states,

    23

  • 8/15/2019 fpaper

    24/27

    is the act of recording “musical impulses.” These musical impulses function not only

    consciously, but subconsciously. Using an algorithm, or even developing an algorithm,

    inherently incorporates musical impulses.

    Will human composers become obsolete? Carson does not think so. Humans havea natural tendency and demand to relate to other humans, and thus have a desire

    to experience compositions produced by human composers. However, as stated, even

    an algorithm expresses some element of a composer’s will and aesthetic. Carson even

    extends the musical impulse to a user merely “pushing play” on some automatic

    composition device.

    He also does not take a post-modern stance of all music being equal. In response

    to Hofstadter, Chopin is still great whether or not his style is replicable by machines.

    Yes, music is a shallow surface structure, but Chopin’s greatness can be attributed

    to his ability to produce surface structures of enormous musical complexity. Car-

    son concludes that Chopin himself is essentially an algorithm of both conscious and

    subconscious processes—but he is a very good algorithm.

    The shallowness of music, according to Carson, is not a fundamental problem.

    Music still has value. People associate with music because it relates to them on a

    multitude of levels, including cultural and emotional levels. Audiences may incorpo-

    rate meaning into their musical experiences, and there is nothing inherently wrong

    with this. However, there is a benefit in increasing musical understanding by dis-

    pelling these myths about composition as some sort of divine, mystical, or spiritual

    process. Knowledge of actual compositional processes should aid audiences in their

    appreciation of music as an art form.

    The question about what EMI was “feeling” when it output the music above is

    perhaps a misguided question. The real question could easily be, “What were David

    Cope’s musical impulses when he wrote EMI, compiled a database of Bach lute suites,

    and pressed the ‘generate’ button?”

    24

  • 8/15/2019 fpaper

    25/27

    5 Conclusion

    Algorithmic composition is still a relatively nascent field given the long course of 

    music history. Its full potential is yet to be explored, and it has yet to completely

    break into the public consciousness. Hofstadter asks:[10]

    Where will we have gotten in twenty years of hard work? In fifty? What

    will be the state of the art in 2084? Who, if anyone, will be able to tell

    “the right stuff” from rubbish? Who will know, who will care, who will

    loudly protest that the last (though tiniest) circle at the center of the

    style-target has still not been reached (and may never be reached)? What

    will such nitpicky details matter, when new Bach and Chopin masterpieces

    applauded by all come gushing out of silicon circuity at a rate faster than

    H2O pours over the edge of Niagara? Will that wondrous new golden age

    of music not be “truly a thing of beauty?”

    And he concludes:

    . . . [T]he day when music is finally and irrevocably reduced to syntacticpattern and pattern alone will be, to my old-fashioned way of looking at

    things, a very dark day indeed.

    The author shares a curiosity in Hofstadter’s questions, but applies a separate

    moral judgment on the future of music. Sharing in Carson’s idea of “musical impulse,”[4]

    the future looks bright for composers to continue to develop systems, as theoreticians

    and composers alike have done throughout history, that will explore new avenues of musical expression and insight. Algorithmic processes do not detract from our hu-

    manity in the slightest. When all is said and done, an algorithm is still a creation

    founded in the human mind, and creations are inherently imbued with the wills of 

    their creators.

    25

  • 8/15/2019 fpaper

    26/27

    References

    [1] Edward Aldwell and Carl Schachter.   Harmony and Voice Leading . Schirmer,

    New York, New York, third edition, 2002.

    [2] Leonard Bernstein.   The Unanswered Question . Harvard University Press, Cam-

    bridge, Massachusetts, 1999.

    [3] Leonard Bernstein.   The Joy of Music . Amadeus Press, LLC, Pompton Plains,

    New Jersey, 2004.

    [4] Benjamin Carson. Personal Interview, December 4, 2009.

    [5] Jacqui Cheng.   Virtual composer makes beautiful music—and stirs 

    controversy .   http://arstechnica.com/science/news/2009/09/

    virtual-composer-makes-beautiful-musicand-stirs-controversy.ars,

    Accessed December 6, 2009.

    [6] David Cope. Computer Models of Musical Creativity . The MIT Press, Cambridge,

    Massachusetts, 2005.

    [7] Peter Elsea.   Fuzzy Logic and Musical Decisions , 1995.   ftp://arts.ucsc.edu/

    pub/ems/FUZZY/Fuzzy_Logic_And_Music.pdf , Accessed December 6, 2009.

    [8] Peter Elsea.  Musical Applications of Fuzzy Logic , 1995.   ftp://arts.ucsc.edu/

    pub/ems/FUZZY/MusAppFuzzy.pdf , Accessed December 6, 2009.

    [9] Robert Gauldin. A Practical Approach to Sixteenth-Century Counterpoint . Wave-

    land Press, Inc., Long Grove, Illinois, 1995.

    [10] Douglas Hofstadter.   Sounds Like Bach .   http://www.unc.edu/~ mumukshu/

    gandhi/gandhi/hofstadter.htm , Accessed December 6, 2009.

    26

  • 8/15/2019 fpaper

    27/27

    [11] Stefan Kostka and Dorothy Payne.   Tonal Harmony, With an Introduction to

    Twentieth-Century Music . McGraw-Hill, 1221 Avenue of the Americas, New

    York, New York, fifth edition, 2004.

    [12] David Lewin.   Generalized Musical Intervals and Transformations . Oxford Uni-

    versity Press, Inc., 198 Madison Avenue, New York, New York, 2007.

    [13] Daniel Morris.   MySong: Automatic Accompaniment for Vocal 

    Melodies , 2008. Online Lecture.   http://academicearth.org/lectures/

     mysong-automatic-accompaniment-for-vocal-melodies, Accessed Decem-

    ber 6, 2009.

    [14] David Nicholls.  American Experimental Music: 1890-1940 . The Press Syndicate

    of the University of Cambridge, New York, New York, 1990.

    [15] Richard Taruskin.  The Oxford History of Western Music , volume I,  The Earliest 

    Notations to the Sixteenth Century . Oxford University Press, Inc., 198 Madison

    Avenue, New York, New York, 2005.

    [16] Richard Taruskin.  The Oxford History of Western Music , volume IV,  The Early Twentieth Century . Oxford University Press, Inc., 198 Madison Avenue, New

    York, New York, 2005.


Recommended