You are on page 1of 77

Chapter 1

Algebra in Music

Music throughout history has generally been played using some type of scale and, at least within the last several centuries, was written down in some form of notation. It helps to be familiar with the basic idea of musical scales and notation in order to more deeply appreciate what composers are doing, what patterns they are using to create their music. The mathematics we introduce in this chapter will clarify the inner coherency of musical scales and some basic composing techniques. This mathematics is just the method of clock arithmetic. Musical scales, key signatures, and these composing techniques are all examples of counting hours and playing with time on a 12-hour clock.

1.1

Musical Scales

In this section we will describe the standard Western musical scales. We shall describe the elementary aspects of such scales and the notation that has been used for Western music for the last few centuries. These are most easily explained using a piano keyboard. Many composers, even when they are composing for other instruments, use a piano to initially sketch out the notes for their compositions. If you do not play an instrument, then it will help to play some of the notes and scales that we will be discussing here. A good approach would be to use software for writing out some of the examples of scores that we look at and playing them. A free version of scoring software can be found at
http://www.musescore.org/

(1.1)

Many of the scores discussed in the book have been written in this software and you can play them after downloading them from the book website. (Or, you can watch the scores being played in video recordings at the book website.)

1.1.1

The C-major scale

The simplest musical scale is the C-major scale. It corresponds to 8 adjacent white keys, an octave, on the piano as shown in Figure 1.1. These white keys are labeled C, D, E, F, G, A, B, C. The last C is considered to be the same note, just an octave higher in pitch. The eight white keys on the piano will play a C-major scale: C D E F G A B C. (1.2)

The black keys are for notes that lie above or below the notes on adjacent white keys. For example, C is a note that is above the note C but below the note D. On the other hand, D is a note that is below the note D but

1. Algebra in Music

above the note C. On the piano, and in our standard (equal-tempered) scale for Western music, the two notes C and D are the same note. We have indicated this by writing C : D above the rst black key on the left side of Figure 1.1. Similar remarks apply to the rest of the black keys, which are used for these accidentals (they actually belong to scales for other musical keys) that occur in music written for the C-major scale. If we were to strike both white and black keys in order from left to right, beginning with the note C on the left, then we would play this scale: C C D D E F F G G A A B C (1.3)

which is called the chromatic scale. The C-major scale is called a diatonic scale. The word diatonic indicates two special tones, and arises from the historical development of this type of scale. These two scales are shown
C : D D : E F : G G : A A : B

FIGURE 1.1. Keys on a piano keyboard for one octave.

in Figure 1.2, along with two other scales that we shall discuss soon. Notice that in Figure 1.1 there are black keys in between some white keys but not others. We have this pattern of how many keys are needed to get from one white key to the next on the piano scale: C D E F G A B C. On the C-major scale, without reference to the piano keyboard, the word semitone is used instead of referring to piano keys. So we can write instead C D E F G A B C. This pattern of key distances or semitones 2 2 1 2 2 2 1 (1.4)
2 semitones 2 semitones 1 semitone 2 semitones 2 semitones 2 semitones 1 semitone 2 keys 2 keys 1 key 2 keys 2 keys 2 keys 1 key

can be used to create major scales with other starting notes.

1.1.2

Other scales

To create a major scale we use the pattern of semitones in (1.4), starting with a different note than C. See Remark 1.1.1 for an explanation of why we will be choosing the fth note in a given scale to begin a new scale. For the C-major scale, the fth note is G. So we shall start with the note G and use the semitone numbers in (1.4) as key distances on the piano keyboard shown in Figure 1.1. This will produce the G-major scale. Starting at G on the keyboard in Figure 1.1, the second note is 2 keys to the right of G, and that second
Mathematics and Music, by James S. Walker.

1.1 Musical Scales

C-major scale

G-major scale

D-major scale

Chromatic Scale

FIGURE 1.2. Some musical scales. Try playing these scales with a real instrument, or a virtual one (using musical scoring software such as M USE S CORE). The notes are scored in the order described in the text. If you are not familiar with elementary musical notation, it will be explained in Section 1.

note is A. Then 2 keys to the right of A is B, and 1 key to the right of B is C. Since we have arrived at the note C, we now wraparound to the starting C on the left.1 From this left hand C we then move 2 more keys to the right to get D. From D we move 2 keys to the right to get E. Then comes a surprise, moving the next 2 keys to the right, gets us to the black key F . Finally, moving 1 key to the right of F gives G. Since we have returned to G, we have completed our G-major scale. Summarizing, the G-major scale is G A B C D E F G. (1.5)

which we got from using this pattern of semitones: G A B C D E F G. Now, suppose we start on the fth note D of this G-major scale. We will produce the scale known as D-major. Using our pattern of semitones as key distances on the keyboard, we get the following results: D E F G A B C D.
wraparound at C 2 keys 2 keys 1 key 2 keys 2 keys 2 keys 1 key 2 semitones 2 semitones 1 semitone 2 semitones 2 semitones 2 semitones 1 semitone

So the D-major scale is D E F G A B C D. (1.6) Try playing these three scales, C-major, G-major, and D-major on a real or virtual instrument to see the similarity in the way the pitches relate. Remark 1.1.1. Why have we emphasized going up a fth? That is, building new major scales based on choosing the fth note of a previous major scale. The reason has to do with the fact that the pitch of a fth note sounds nearly as harmonious as an octave change. A note that is an octave higher sounds so close in tone to the note that is an octave lower that we call them the same note. In fact, males and females generally sing about an octave apart, so this may be the reason that we hear notes an octave apart as the same (due to long historical practice of males and females singing together). A fth note on a musical scale is nearly as harmonious with the initial note as the note an octave higher. So when people sing together, as they have done throughout human
1 On the piano keyboard itself we could just continue to the right, but we want to explain our scales in a way that applies to other instruments besides the piano, so we shall perform this wraparound. There is another method, which we will describe shortly, that does not use the piano keyboard at all.

Mathematics and Music, by James S. Walker.

1. Algebra in Music

life, it is extremely important to have different musical scales that are generated by going up in fths. In fact, studies have been reported where people were asked to sing an octave higher note and a great many people actually sang a note that was only a fth higher. This is not just a Western cultural bias. Throughout the world, almost all musical scales contain two notes that are an octave apart and a middle note that corresponds to the fth note in our classical Western scales. There is a physical basis for this predominance of octaves and fths which we will describe in the next chapter when we take up the relation between musical pitch and frequency of air vibrations in musical sound.

1.1.3

Scales and Clock Arithmetic

The method of using the piano keyboard for describing scales is workable, but a bit cumbersome. There is an elegant mathematical way of creating scales using the notion of clock arithmetic (or arithmetic modulo 12 to use the more technical terminology). In our discussion above, we wrapped around on the piano keyboard shown in Figure 1.1 when arriving at the higher octave C on the right side, so that we would be back to the lower octave C on the left side. The easiest way to express this wraparound property is to put the notes from the chromatic scale onto the hour positions on a clock face as shown in Figure 1.3. C B A : B C : D D

D : E

G : A G F : G F

FIGURE 1.3. The Chromatic Clock. The chromatic scale arranged on a clock face (starting at 0).

This clock has the top hour start at 0 rather than 12. Using this clock gure, which we shall call the Chromatic Clock, we can create major scales using the sequence 2 2 1 2 2 2 1 (1.7) as hours to add in moving around the hour positions on the clock. We agree, however, that when we circle around the top of the clock (the number 0) we will always reset to 0 by subtracting 12 (just as in telling time with this type of clock). Example 1.1.1 (C-major scale). If we start at the note C for our C-major scale at hour 0 and then apply the sequence in (1.7) as hour moves around the clock, then we get the C-major scale. The note C is at hour 0, so adding 2 to 0 gives 2 and note D is at 2 hours, adding 2 to 2 gives 4 and note E is at hour 4. We then add 1 to 4 to get 5, and note F is at hour 5, adding 2 to 5 gives 7 and note G is at hour 7, adding 2 to 7 gives 9 and note A is at hour 9, then adding 2 to 9 gives 11 and note B is at hour 11. Finally, adding 1 to 11 gives 12 and we are
Mathematics and Music, by James S. Walker.

1.1 Musical Scales

at the top of the clock, so we subtract 12 and get 0, and note C is at hour 0. We have found the C-major scale: C D E F G A B C. (1.8)

We can summarize our calculations with this diagram: 0 C


+2

2 D

+2

4 E

5 F

+1

+2

7 G

+2

9 A

11
12 at top

+2

+1

0 C.

(1.9)

Example 1.1.2 (G-major scale). More interestingly, if we start at G instead (to get a G-major scale) then we are starting at hour 7 for G. Adding 2 hours gives 9, which is A. Then adding 2 to 9 gives 11, which is B. Adding 1 to 11 gives 12 but we are at the top of the clock so we subtract 12 to get 0, which gives C. Now, adding 2 to 0 gives 2, which is D. Adding 2 to 2 gives 4, which is E. Adding 2 to 4 is 6, which is F . Finally, adding 1 to 6 gives 7, which is G, our starting note. Thus, we have obtained the G-major scale: G A B C D E F G. (1.10)

We can summarize our calculations with this diagram: 7 G


+2

9 A

11
12 at top

+2

+1

0 C

+2

2 D

+2

4 E

+2

6 F

+1

7 G.

Example 1.1.3 (D-major scale). We can also use this method to produce the D-major scale. The note D has hour number 2, and the calculations go as follows: 2 D
+2

4 E

+2

6 F

+1

7 G

+2

9 A

11
12 at top

+2

+2

1 C

+1

2 D

to get the D-major scale: D E F G A B C D. (1.11)

Example 1.1.4 (A-major scale). Finally, lets go up a fth from the D-major scale. We will start from its fth note, A, and create the A-major scale. The hour number for A is 9 and the calculations go as follows: 9 A 11
12 at top +2 +2

1 C

+1

2 D

+2

4 E

+2

6 F

+2

8 G

+1

9 A

to obtain the A-major scale: A B C D E F G A. (1.12)

Remark 1.1.2. Looking at these successive scales, we see that every time we went up a fth we found the same notes as on the previous scale, except for one, and we picked up an extra sharp note in the scale. This phenomenon is summarized in the famous Circle of Fifths which we shall describe later.

Exercises
1.1.1. Use the clock arithmetic method to go up a fth from the A-major scale, shown in (1.12), to nd the E-major scale.
Mathematics and Music, by James S. Walker.

1. Algebra in Music

1.1.2. Explain why going up a fth note in a scale is always done by adding 7 hours on the Chromatic Clock. 1.1.3. Suppose you go down a fth on the C-major scale, shown in (1.2). That is, you take as your starting note the fth note down from the higher octave C note, which is the note F. Use the clock arithmetic method to nd the F-major scale, which starts with this note F. [Note: Make sure you do not use a note letter more than once, resulting in one atted note.] 1.1.4. Starting with the note F , use this sequence of additions on the Chromatic Clock: +2 +2 +3 +2 +3

to generate a scale of six notes. (Notice that its ve distinct notes are the 5 black keys on a piano.) 1.1.5 (Harmonic minor scales). An harmonic minor scale is obtained from a given starting note by using the following sequence of adds in our clock arithmetic method: +2 +1 +2 +2 +1 +3 + 1.

Use these additions to generate an harmonic scale that begins with C (the harmonic C-minor scale). (Remember to not repeat note letters, this will determine whether you use ats or sharps for accidentals.) 1.1.6. Parts (a) through (e) of this exercise develop the reasoning behind why building new major scales by going up a fth will share all but one of the same notes as on the previous scale, and typically pick up an extra sharp. (a) Show that when we start at the fth note of a major scale, the sequence of semitone changes (clock adds) for the scale looks like this: +2 +2 +1 +2 +2 +1 + 2. (b) By comparing the sequence of clock arithmetic adds in part (a) to the one we have used to build new major scales, explain why a new major scale will share all of its notes with the original major scale, except for one, and the note that differs will occur one hour back from the end note. (c) Starting with the C-major scale, explain why the scale going up by a fth (the G-major scale) will share the same notes but one with the C-major scale and pick up one sharp note. (d) Now, starting with the G-major scale, explain why the scale going up by a fth (the D-major scale) will share the same notes but one with the G-major scale and pick up one sharp note (so it has two sharps). (e) Show that if we keep going up a fth to create a new major scale, the pattern of picking up one more sharp will continue. Until we reach the scale that begins with which note? 1.1.7 (Natural minor scales). The natural A-minor scale is obtained from the C-major scale by starting at its sixth note, A, and then using the following sequence of adds in our clock arithmetic method: +2 +1 +2 +2 +1 +2 + 2. (1.13)

(a) Using our clock arithmetic method with this sequence of adding hours, starting at the note A on the clock (see Figure 1.3), show that the following scale is produced A which is called the A-minor scale. (b) Why does the A-minor scale contain exactly the same notes as the C-major scale? (c) Go up a fth on the A-minor scale and use the clock arithmetic method with the sequence in (1.13) to produce a new minor scale starting at E, the E-minor scale. (d) Go up a fth on the E-minor scale and use the clock arithmetic method with the sequence in (1.13) to produce a new minor scale starting at B, the B-minor scale.
Mathematics and Music, by James S. Walker.

(1.14)

1.2 Elementary Musical Notation

(e) Why does each successive minor scale obtained by going up a fth on the previous minor scale have the property that it shares the same notes but one? Also, why generally is one additional sharp added (although there are exceptions)? Hint: This generalizes Exercise 1.1.6 to the case of minor scales. 1.1.8. Denote the set of hours on the Chromatic Clock by H, so H = {0, 1, 2, 3, . . . , 10, 11}. Show that addition of hours in H satises these four properties: (1) For each j, k in H, the sum j + k is in H (provided 12 is subtracted when passing the top of the Chromatic Clock). (2) For each j, k, in H, we have (j + k) + = j + (k + ). (3) For each j in H, we have 0 + j = j and j + 0 = j. (4) For each j in H, there is a unique k in H that satises j + k = 0 and k + j = 0. Remark 1.1.3. Exercise 1.1.8 shows that the set of hours H on the Chromatic Clock is a mathematical group. Mathematical groups are one of the most fundamental algebraic structures in mathematics.

1.2

Elementary Musical Notation

The purpose of this section is to briey discuss some elementary musical notation. Please feel free to quickly skim, or even skip, this section if you already know how notes are scored and what time signatures are. The treble clef is the staff marked with the G-clef shown on the left side of Figure 1.4. On the right side of

F E D C B A G F E

FIGURE 1.4. Left: The treble clef. Right: The treble clef with note positions marked.

Figure 1.4 we have indicated where the notes are marked on this clef. The notes that are marked on the lines are E, G, B, D, F. A mnemonic device for remembering that sequence is Every Good Boy Does Fine or, as is more commonly said in Canada and England, Every Good Boy Deserves Favour.2 The notes in the spaces can be remembered with another mnemonic, this time a sort of acronym: FACE. As an example of this notation, we show some sequences of notes on treble clefs in Figure 1.5.
2

There was an album with that title, Every Good Boy Deserves Favour, by The Moody Blues produced in the 1960s.
Mathematics and Music, by James S. Walker.

1. Algebra in Music

The notes D, G, A, G, B, D, E, D

The C-major scale C, D, E, F, G, A, B, C

FIGURE 1.5. Left: A sequence of notes on a treble clef. Right: The C-major scale on a treble clef.

A second clef is often used together with the treble clef, the bass clef. The bass clef contains lower pitch notes than the treble clef, as its name indicates. In Figure 1.6 we show the bass clef and the positions where notes are marked on it. A mnemonic for the notes marked on the lines is Good Boys Do Fine Always and a mnemonic for the notes marked between the lines is All Cows Eat Grass.

A G F E D C B A G

FIGURE 1.6. Left: The bass clef. Right: The bass clef with note positions marked.

Finally, these two clefs can be combined to create the grand staff. In Figure 1.7 we show the grand staff and the positions where notes are marked on it. Notice, in particular, that the middle note between the two clefs is C. It is referred to as middle C.
F E D BC GA F E D C B A F G DE BC A G

FIGURE 1.7. Left: The grand staff. Right: The grand staff with note positions marked.

1.2.1

Time Signatures

In Western music, as commonly practiced, there is an intricate division of the length of time that notes are
Mathematics and Music, by James S. Walker.

1.2 Elementary Musical Notation

played. This division of time is part of the rhythm of the music and is indicated in musical scores by time signatures. For now, we will only describe the two most common time signatures: the 4 time signature and 4 the 3 time signature. 4 The
4 4

time signature

The notation 4 looks like a fraction, but without the fraction bar. We shall see that there is some relation 4 between the time signature, 4 , and the fraction, 4 . 4 4 The top number in the time signature, in this case a 4, indicates that there will be four fundamental beats (notes or rests) in each measure. While the bottom number, also a 4, indicates that each of those fundamental beats will be quarter notes (or rests of duration equal to a quarter note). Here is an example:
1 whole note 2 half notes 4 quarter notes 8 eighth notes 16 sixteenth notes

1=

4 4

1 2

1 2

4 4

1 4

4 4

1 8

4 4

16

1 16

4 4

In the rst measure, there is just one note, a whole note. One whole note equals 4 quarter notes, so we have the 1 4 equation 1 = 4 . While in the second measure, we have 2 half notes, so that gives the equation 2 + 1 = 4 . The 4 2 4 1 third measure consists of four quarter notes, and that gives the equation 4 4 = 4 . The rest of the measures, and the corresponding equations with fractions, should be clear at this point. In each case we have successively divided the length of a note by 2 and formed the new measure by twice as many repetitions of this halved note. In every case, this leads to an equation with fractions that has 4 on its right side. The connection of 4 with the 4 4 time signature 4 should now be clear. 4 As one more example of this division of note lengths by 2, we show one measure consisting of 32 repetitions of a 1/32 note (each note being of duration 1/32nd of the duration of the whole measure):
32 thirty-secondth notes: 32
1 32

4 4

In addition to beats consisting of notes, there are also beats consisting of rests, which are pauses between notes. In 4 time, the fundamental rest is a quarter rest. Here is an example that illustrates both notes and rests: 4
1 whole rest 1 half note, 1 half rest notes separated by
1 4

rest,

1 8

rest,

1 16

rest,

1 32

rest

1=

4 4

1 2

1 2

4 4

1 4

1 4

1 8

1 8

1 16

1 16

1 32

1 32

2 32

4 4

4 In the rst measure, one whole rest takes up the whole measure. This gives the equation 1 = 4 . In the second 1 measure, there is one half note followed by one half rest, resulting in the equation 1 + 2 = 4 . Finally, in the 2 4 third measure, there is a more complex sequence of notes and rests obtained by further division of duration by 2. There is a quarter note, followed by a quarter rest, and then an eighth note followed by an eighth rest, and so on. The sum of all the beat durations in the measure produces the equation:

1 1 1 1 1 1 1 1 1 4 + + + + + + + +2 = . 4 4 8 8 16 16 32 32 32 4
Mathematics and Music, by James S. Walker.

10 The nal aspects of of these notations:


`1
2

1. Algebra in Music
4 4

time that we will discuss here are dotted notes, ties, and slurs. Here is an example
1 4 4 4

1 4

`1
4

1 8

+2

1 8

1 4

1 8

4 4

1 8

+2

1 4

+4

1 16

4 4

In the rst measure, there is a dotted half note. The duration of any dotted note is 1.5 times the duration of the note symbol without the dot. So the dotted half note has duration 1.5 (1/2), which we will express as 1 1 1 1 4 1 2 + 4 . This gives the equation for the duration of the beats in the rst measure to be 2 + 4 + 4 = 4 . In the second measure, there is a dotted quarter note, followed by two eighth rests, and then a quarter note tied 1 1 4 together with an eighth note. The resulting duration equation is 1 + 8 + 2 1 + 4 + 1 = 4 . The tying of the 4 8 8 notes has no effect in calculating their duration. The tied quarter note and eighth note have the same duration, 1 1 4 + 8 , as they would if they occurred separately without a tie. The tie indicates how the notes are to be played together, without an audible silence between them. Finally, the third measure has two eighth notes (connected with a slur), followed by a quarter rest and a quarter note, followed by four sixteenth notes (connected with a slur). The slurs have no effect on the durations, as indicated by the equation shown for the third measure. They indicate, as their name implies, that the notes should be slurred together as they are played so that no silences are audible between the individual notes. Before we leave the topic of the 4 time signature, we should point out that it is the most common time 4 signature. Consequently, when a score does not have a time signature given, then the convention is that the time signature is 4 . For example, we have this music notation equation: 4

=
There are some additional aspects to the when they appear in later examples. The
3 4 4 4

time signature. We will discuss them in the exercises, or as needed

time signature

The other time signature that is frequently employed is the 3 time signature. The basic ideas are the same as 4 with 4 , so we shall just briey discuss this case. The basic beat-length is 1/4, either a quarter note or a quarter 4 rest, and there are 3 beats to a measure. Here is an example:
3
1 4

3 4

`1

1 8

1 8

`1

1 16

1 32

1 32

3 4

1 16

1 8

1 16

1 4

1 8

1 16

1 16

3 4

In the rst measure, there is a quarter note followed by a quarter rest and a quarter note. Since there are 1 3 beats in a row, we get the equation 3 4 = 3 . The second measure begins with a dotted quarter note, 4 1 expressed as 1 + 8 , followed by an eighth rest, a dotted eighth note, and nishing with a 1/32 rest and 4 1 1 1 1 3 1/32 note. This gives the equation 1 + 1 + 1 + 8 + 16 + 32 + 32 = 4 for the measure. Again we can 4 8 8
Mathematics and Music, by James S. Walker.

1.2 Elementary Musical Notation

11

see the connection of time signature notation with fractions, noticing that the sum of the beat durations for 3 each measure produces the fraction 4 for this time signature of 3 . The nal measure in the example above 4 provides further conrmation of this connection. It consists of a sixteenth rest, followed by an eighth note and 1 1 a sixteenth rest, followed by a quarter rest (producing durations of 16 + 1 + 16 + 1 ), and nishing with an 8 4 1 1 1 eighth note followed by a sixteenth note and sixteenth rest ( 8 + 16 + 16 ). Altogether these durations produce 1 1 1 1 1 1 the equation: 16 + 8 + 16 + 1 + 8 + 16 + 16 = 3 . 4 4

Exercises
1.2.1. For the following treble clef

nd the notes:

1.2.2. For the following bass clef

nd the notes:

1.2.3. For the following Grand Staff

nd the notes: Treble Clef:

Bass Clef:
Mathematics and Music, by James S. Walker.

12
1.2.4. For the following Grand Staff

1. Algebra in Music

nd the notes: Treble Clef:

Bass Clef:

1.2.5 (Alto clef). The Alto Clef is another clef that is sometimes used. Here is an Alto Clef, where we have indicated on the left side where middle C is located:

Find the notes marked on this Alto Clef:

1.2.6 (Tenor clef). The Tenor Clef is another clef that is sometimes used. Here is a Tenor Clef, where we have indicated on the left side where middle C is located:

Find the notes marked on this Tenor Clef:

1.2.7. The following score uses all of the clefs that we have discussed. Name the notes that are scored.

Treble Clef: Tenor Clef:

Alto Clef: Bass Clef:


Mathematics and Music, by James S. Walker.

1.2 Elementary Musical Notation

13

1.2.8. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 4 time signature: 4

1.2.9. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 4 time signature: 4

1.2.10. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 3 time signature: 4

1.2.11. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 3 time signature: 4

Tuplets Another aspect of rhythm in time signatures is the grouping of notes whose duration is not a power of 2 division of a beat length. These groupings are usually called tuplets. The most common tuplet is a triplet, which has 3 notes in a group. Here is an example of four triplets in 3 time: 4
triplet: 1 note-length 4 triplet: 1 note-length 8
1 triplet: 16 note-length 1 triplet: 4 note-length

`1
4

1 4

1 4

3 4

1 8

`1
8

1 4

+2

1 8

3 4

1 8

1 16

1 16

1 4

`1
4

3 4

The rst measure begins with a triplet of 3 eighth notes. Although the triplet is notated with eighth notes, their durations 1 1 add up to the duration of a quarter note, so each of the notes actually has a duration of 3 1 = 12 . In the equation below 4 1 the measure, we have listed the total duration of the triplet as 4 . Likewise, in the second measure, there is a triplet
Mathematics and Music, by James S. Walker.

14

1. Algebra in Music

denoted using sixteenth notes. But the duration of the entire triplet is equal to the duration of an eighth note, which we have denoted by 1 in the equation below the measure. The convention is that the duration of the triplet is twice as 8 long as the duration of the note that is repeated to form the triplet. For example, in the third measure there is a triplet of 1 1/32nd notes, so the length of the whole triplet is 1/16 which we have written as 16 in the equation for the measure. Based on the discussion just given, please solve the following 4 exercises. 1.2.12. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 3 time signature: 4

For each triplet, what are the actual durations of the notes? 1.2.13. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 4 time signature: 4

For each triplet, what are the actual durations of the notes? 1.2.14. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 3 time signature: 4

For each tuplet, what are the actual durations of the notes? 1.2.15. For the following score, for each measure write the equation with fractions that corresponds to the note lengths, given the 4 time signature: 4

For each tuplet, what are the actual durations of the notes? Tempo An aspect related to how the rhythm of notes is played is its overall speed, its tempo. With the invention of the metronome a tempo could be specied very accurately. Often in scores a metronome marking is given for the tempo. For example, when the following marking
Mathematics and Music, by James S. Walker.

1.2 Elementary Musical Notation

15

appears at the beginning of the score, then the tempo is set at 120 beats/minute (bpm). In this case the beat is a quarter note and 120 quarter notes will have a duration of 1 minute. Therefore, the duration of a quarter note is 0.5 seconds. Based on this discussion, solve the following two exercises. 1.2.16. For each of the tempo marks below, nd the length in seconds of a quarter note.

1.2.17. Some musical scores use Italian names to indicate tempo, such as Adagio or Allegro or Andante. Precise bpp values are not always specied, a metronomically precise tempo being up to the performer(s) and/or conductor. Suppose the performance is to be played at tempos that use these values for bpm (where a beat is a quarter note): Adagio = 40 Allegro = 120 Andante = 80

In each case, nd the length in seconds of a quarter note. Automatic Music Scoring In computer music scoring software, such as M USE S CORE, when a note is entered the rests remaining in the measure are automatically computed. Here is an example:
Add 1/32nd note

The next two exercises deal with the algebra underlying how the software calculates the remaining rests in a measure after entering a note. 1.2.18. In the example above, let x stand for the duration left in the measure after the 1/32nd note is entered. Show that the following equation holds 1 3 +x= 32 4 and solve for x. Show that x can be expressed in a way that reects the rests shown in the example above. 1.2.19. Use the method outlined in the previous exercise to nd the sequence of rests that remain in the measure after entering the notes indicated in each of these three cases: a. Add a dotted sixteenth note:

b. Add a dotted quarter note (right after the sixteenth note):

Mathematics and Music, by James S. Walker.

16
c. Add a dotted sixteenth note:

1. Algebra in Music

1.3

Key Signatures and The Circle of Fifths

The musical examples given in the next section are written in specic keys. In this section we will describe the notation for these keys and how they are connected mathematically by the famous Circle of Fifths. The Circle of Fifths is derived using the mathematics of addition on the Chromatic Clock. For each of the examples in the next section, a specic scale is used for the vast majority of the notes. This emphasized scale is called the key for the music and is indicated on the scale by marking the sharps, or ats, that occur on the scales notes. For example, if a score is written using predominantly notes from the scale of A-major, then the score will begin with 3 sharp symbols marked on the lines for the notes C, F, and G. See the top of Figure 1.8. As shown in (1.12), these notes appear as the sharped notes C , F , and G in the A-major scale. By placing sharps at the positions for C, F, and G (notice that the sharp for G is placed in the space above the F-line, which is a G-note space), it is then unnecessary to mark them on each appearance of one of the three sharped notes in the A-major scale. For example, in the top of Figure 1.8, we show a sequence of notes containing several instances of C , F , and G , but in the score these notes are not marked with sharps. The key signature of 3 sharps at the start, indicates that they are sharped notes. Moreover, because of the properties of the celebrated Circle of Fifths, it turns out that just marking the 3 ats is sufcient to identify the A-major scale. When a note is used that is not part of the A-major scale, like C , then it is specically marked in the score. We have illustrated this in the top of Figure 1.8. The C-major scale has no ats or sharps. So when no ats or sharps are written at the start, as shown in the bottom of Figure 1.8, then that indicates the C-major scale. All notes in the score are naturals, no ats or sharps, unless the notes are specically marked in the score. We show an example of such marked notes in the bottom of Figure 1.8. Notice that the last note is marked as G , which indicates that that note is G and is not to be played sharp or at.

A-major

C C-major

FIGURE 1.8. Top: A sequence of notes in the key of A-major. Bottom: A sequence of notes in the key of C-major.
Mathematics and Music, by James S. Walker.

1.3 Key Signatures and The Circle of Fifths

17

1.3.1

Circle of Fifths

Key signatures are connected in an organized way. This organization is called the Circle of Fifths. The Circle of Fifths for the key signatures of the major scales is shown in Figure 1.9.

C F 1 B 2 0 G 1 2 D

4 5 D 6 or 6 G or F 5

4 B

FIGURE 1.9. The Circle of Fifths. Key signatures for the major scales.

This gure should be interpreted as follows. Starting at the top, the position 0 for the key of C-major, each time we move one hourly position clockwise (numbered by sharps), we reach a new key that shares the same notes as the previous key except for one new note. Each new key can be indicated on the clef with one more than the preceding key on the Circle of Fifths. This occurs until we reach the position indicated by 6 or 6 marking the key of F , which has 6 sharps, and is also called the key of G , which has 6 ats. On the other hand, if we start at position 0 for the key of C-major, then each time we move one hourly position counterclockwise (numbered by ats), we reach a new key that shares the same notes as the previous key except for one new note, and this new key can be indicated on the clef with one more than the preceding key on the Circle of Fifths. The reason why the Circle of Fifths can organize key signatures in this fashion was described in Exercise 1.1.6 above. We will explain the solution of that exercise here. Suppose we begin with the key of C-major, which has 0 sharps and 0 ats, at the top of the Circle of Fifths. The scale for C-major is constructed from the Chromatic Clock in Figure 1.3 using the following calculations: 0 C
+2

2 D

+2

4 E

5 F

+1

+2

7 G

+2

9 A

11
12 at top

+2

+1

0 C.

When we go up a fth to the next key of G-major at position 1 on the Circle of Fifths, then we are performing these calculations on the Chromatic Clock starting at the note G:
Mathematics and Music, by James S. Walker.

18

1. Algebra in Music
matches end sequence for previous key match at beginning new note, 1 hour before end

7 G

+2

9 A

11
12 at top

+2

+1

0 C

+2

2 D

+2

4 E

+2

6 F

+1

7 G.

This diagram shows why the key of G-major, which is up a fth from C-major, will have one new note that differs from C-major. And that new note will be the one that immediately precedes the end note on the Chromatic Clock. Since the note immediately preceding the end note of G on the Chromatic Clock is F , the key of G-major must have this one sharped note. Notice that at position 1 on the Circle of Fifths in Figure 1.9 we have one sharp in the key signature, and it is marking the note F on the treble clef. Similar reasoning applies each time we move another hourly position clockwise through positions 1 to 6 on the Circle of Fifths, which corresponds to going up a fth on the Chromatic Clock (adding 7 hours on the Chromatic Clock). We always end up with one note that is different, the one preceding the end, and that note preceding the end is a sharp note (at least until we get to position 6 on the Circle of Fifths). When we get to position 6 on the Circle of Fifths, we have reached F which has the note F preceding it. So it appears as if we did not pick up an extra sharp. We have the notes F G A B C D F F

In musical notation, however, it is customary to never use a note letter more than once in a scale. Since the convention in music is that putting a sharp on a note raises that note up one semitone, we can write F as E . The notes are said to be enharmonic.3 Therefore, the scale above is typically expressed as F G A B C D E F

This scale has 7 sharps, but one of them is a repeat. So the key signature for this scale has just 6 sharps. Going counterclockwise on the Circle of Fifths will produce new key signatures by adding ats. Each additional hourly position counterclockwise will add a at to the previous key signature. When we go counterclockwise on the Circle of Fifths we are going down a fth from the preceding key. For example, if we start again at C-major at hour 0 on the Circle of Fifths, then we have these calculations on the Chromatic Clock: 0 C
+2

2 D

+2

4 E

5 F

+1

+2

7 G

+2

9 A

11
12 at top

+2

+1

0 C.

However, when we go down a fth to the next key of F-major at position 1 on the Circle of Fifths, then we are performing these calculations on the Chromatic Clock, starting at the note F:
matches previous key
+2 +2

new note (2 before C)


+1 +2

matches previous keys beginning


+2 +2 +1

5 7 9 10 0 2 4 5 F G A B C D E F. This diagram shows why the key of F-major, which is down a fth from C-major, will have one new note that differs from C-major and that note will be the one that is back 2 hours from C on the Chromatic Clock. Since the note that is 2 hours back from C on the Chromatic Clock is B , the key of F-major must have one at. Notice that at position 1 on the Circle of Fifths in Figure 1.9 we have one at and it is marking the note B on the treble scale. As we continue to move counterclockwise around the Circle of Fifths, we pick up one new atted note at a time. Until we reach the position 6, meeting up with the key of F -major, which is
3

Likewise, putting a at on a note lowers that note by a semitone. Including unusual cases, like C , which is enharmonic with B.
Mathematics and Music, by James S. Walker.

1.3 Key Signatures and The Circle of Fifths

19

enharmonic with the key of G -major. Each note of F -major is enharmonic with the corresponding note of G -major.

Exercises
1.3.1 (7 sharps or 7 ats). Occasionally, especially in classical music, one encounters key signatures with either 7 sharps or 7 ats. Place those two key signatures on the Circle of Fifths, and determine their names in both major and minor keys. 1.3.2 (Circle of Fifths for Minor Keys). In Figure 1.10 we show the Circle of Fifths for the natural minor keys. In this exercise, we outline the mathematical explanation for the structure of this Circle of Fifths. Recall that the natural A-minor key, denoted as a in Figure 1.10, was obtained from the C-major key by going up a sixth (i.e. reading off the notes from the C-major scale starting at the sixth note A), so it has the same notes as the C-major scale. Each minor scale is obtained from a major scale in the same way, going up a sixth from the major scales starting note. Therefore, to get the minor keys to arrange in a Circle of Fifths, we just have to successively go up a fth from A-minor, or successively down a fth from A-minor. Your exercise is to check that doing this produces the keys shown in Figure 1.10. [Hint: remember that going up a fth involves adding 7 hours on the Chromatic Clock, and going down a fth involves subtracting 7 hours.]

a d 1 g 2 0 e 1 2 b

4 5 b 6 or 6 e or d 5

4 g

FIGURE 1.10. The Circle of Fifths for the natural minor keys. 1.3.3 (Chord structure). In this gure we show the fundamental triadic (3-note) chords in the key of C-major:

I C

ii D

iii E

IV F

V G

vi A

vii B

I C

The major chords are labeled by upper case Roman numerals, I, IV, V, and are also labeled by their base notes (C, F, and
Mathematics and Music, by James S. Walker.

20

1. Algebra in Music

G)for instance, chord V is also called a G-major chord. The minor chords are labeled by lower case Roman numerals, ii, iii, vi, vii , and are labeled by their base notes (D, E, A, B)for instance, chord ii is also called a D-minor chord. [The superscript is indicating a diminished chord, which will be explained in this exercise.] (a) Find the notes that comprise each of these chords. (b) Mark off the positions of each of the notes in these chords on the Chromatic Clock. What pattern do you see in the intervals (number of semitones) between the notes in the chords for the major chords, I, IV, and V, and for the minor chords, ii, iii, vi? What is different about the diminished minor chord vii ? (c) If the same pattern of intervals (number of semitones) had been used for the base note B in the diminished minor chord vii as was used for the other minor chords, show that the notes in the chord would have been B, D, F . Show that that chord is the minor chord iii in the key of G-major, which is the next adjacent key moving clockwise on the Circle of Fifths. 1.3.4 (Pentatonic scale). The most common pentatonic scale (5 distinct notes) is the following: C D E G A C. (1.15)

which is often referred to as the Pentatonic C-major scale. (a) Mark the notes of the Pentatonic-C scale on the Chromatic Clock, and show that the intervals between its notes follow this pattern: 2 2 3 2 3. (1.16) (b) In his opera, Turandot, Puccini uses some Chinese melodies. Written in Western notation, one of these Chinese melodies is

Write out the sequence of notes used for this melody. (c) For the sequence of notes in part (b), you should have only used letters from this 5 member set: {A, B, D, E, G}. The melody appears to be based on a pentatonic (5-note) scale, but the notes are not exactly the same as for the Pentatonic-C scale (since they include B and do not include C). Show that the ve notes that are used for the melody are another pentatonic scale obtained from the Pentatonic-C scale by addition on the Chromatic Clock, and that the intervals between this pentatonic scales notes follow the same pattern as in (1.16). What would be a good name for this pentatonic scale? 1.3.5 (Chords in Pentatonic Scales?). Generally, music that is written in pentatonic scales does not employ chords. This exercise will provide some explanation for this. Suppose that we are using the Pentatonic-C scale in (1.15). Find the locations of the chord C-E-G on the Chromatic Clock, and the intervals separating these notes. Can you nd a three-note chord, using D as the base note, which has the same (or reversed) intervals? What about a three-note chord that has E as base note?

1.4

Diatonic Transformations Scale Shifts

A common technique in musical composition is to use transformations of small groups of notes. The most common types of transformations are known as diatonic scale transformations. Diatonic scale transformations that composers frequently use are (1) diatonic scale shifts, (2) diatonic scale inversions, (3) palindrome symmetries. In this section, we will describe diatonic scale shifts, and take up the other two types in the next section.
Mathematics and Music, by James S. Walker.

1.4 Diatonic Transformations Scale Shifts


To see how diatonic scale shifts work, lets consider the C-major scale: C D E F G A B C

21

A diatonic scale shift is done by shifting a set of notes up or down on the scale. For example, suppose we have four notes: E G F D. If we shift each of them up 1 degree on the scale, then we get the notes F A G E. In Figure 1.11, we show how this diatonic scale shift, which we denote by S1 , appears on a score. It is easy to see how the shifting up by 1 affects the notes in the score.

S1

FIGURE 1.11. A diatonic scale shift of four notes.

We could also have shifted the notes E G F D down by 1 degree on the scale, obtaining D F E C. This diatonic scale shift is denoted S1 . If the notes are shifted by 0 degrees on the scale, then they do not change, so the sequence of notes is repeated. Repetition is very important in music, so the diatonic scale shift S0 does appear frequently. In Figure 1.12 we show the diatonic scale shift S0 , as well as the diatonic scale shifts S1 and S3 , acting on a set of 3 notes.

E
S0

E
S1

E
S3

FIGURE 1.12. A sequence of diatonic scale shifts of three notes.

Example 1.4.1 (Diatonic scale shifts in chords). Diatonic scale shifts occur in chord progressions (changing from one chord to another). For instance, in the gure below we show the basic triadic chords in the key of C-major, and the fact that the diatonic scale shift S1 is used to change from one chord to the next.

S1

E ii

S1

E iii

S1

E IV S1 V E

S1

E vi

S1

E vii

S1

EI

Some standard chord progressions can be written in terms of diatonic scale shifts. For example, the proS S gression I-IV-V can be written as I 3 IV 1 V. Another case is the progression I-vi-ii-V, which can be written as I 5 vi ii 3 V. Example 1.4.2 (Chord progressions in Pachebels Canon in D). An example of sequences of diatonic scale shifts in a famous piece of classical music is Pachebels Canon in D. In Figure 1.13 we show the chords that describe the main theme from this piece. There is a sequence of several diatonic scale shifts that link these chords.
Mathematics and Music, by James S. Walker. S S4 S

22
S2

1. Algebra in Music
S2

T
S2 S2

T
S0

T
S1

FIGURE 1.13. Sequencing of diatonic scale shifts in the theme of Pachebels Canon in D.

Combining a succession of diatonic scale shifts in this way is known as sequencing in music theory. The whole succession is called a tonal sequence.4 Diatonic scale shifts can also be applied to short sections of a melody, and not only to chords. We shall now examine some sequences of diatonic scale shifts of this kind, starting with some examples from classical music. Example 1.4.3 (Beethovens Fifth Symphony). A famous example of using diatonic scale shifts occurs in the beginning of Beethovens Fifth Symphony. In Figure 1.14, we show the opening measures of the score with the diatonic scale shifts marked. The tonal sequencing that Beethoven performs on the initial four note motive is remarkable for the power it produces in the sound of this symphony.

S1

S6

S1

S1

FIGURE 1.14. Opening measures of Beethovens Symphony Number 5. To hear a performance of this music, go to Videos link at book website and select Introduction to Beethovens Fifth Symphony.

Example 1.4.4 (J.S. Bachs Inventio 11). J.S. Bach used diatonic scale shifts frequently as well. In Figure 1.15 we show the opening measures of Bachs Inventio 11 (BWV 782) with four different diatonic scale shifts marked (there are several others as well). Notice that in the last two diatonic scale shifts, the ones labelled S1 and S2 , there are various accidentals for the notes. That is, some of the notes are marked with a , or , or . For these diatonic scale shifts, the sequences are referred to as modied sequences in music theory.5 To indicate that the diatonic scale shifts have been modied, we have marked them with a prime symbol (e.g. S1 rather than S1 ). Example 1.4.5 (Alicia Keys If I Aint Got You). Besides classical music, diatonic scale shifts occur in popular music as well. In Figure 1.16 we have shown the opening measures of the song, If I Aint Got You, by Alicia Keys. The diatonic scale shifts in this song are quite interesting in that they have a hierarchical structure. The hierarchy is organized into at least two levels according to a short timescale and a long timescale.
4 S. Kostka and D. Payne, Tonal Harmony, with an introduction to twentieth-century music, Sixth Edition, McGraw-Hill, 2009, p. 103. 5 S. Kostka and D. Payne, Tonal Harmony, with an introduction to twentieth-century music, Sixth Edition, McGraw-Hill, 2009, p. 104.

Mathematics and Music, by James S. Walker.

1.4 Diatonic Transformations Scale Shifts


S1

23
S1

S1

S2

FIGURE 1.15. Opening measures of Bachs Inventio 11.

FIGURE 1.16. Opening measures of Keys If I Aint Got You.

For example, in the rst measure there is a three note set (motive) consisting of this triplet:

(1.17)
Mathematics and Music, by James S. Walker.

24

1. Algebra in Music

to which the diatonic scale shift S0 is applied two times to get the entire measure:
S0

S0

The next two measures of the treble clef are obtained by rst applying S0 to the rst measure, which is a longer timescale application of S0 , and then applying S1 to the second measure. This produces the rst 3 measures:
S0

S1

Notice that the diatonic scale shifts S0 and S1 are also applied to the single notes in each measure on the bass clef. In fact, you can check that the diatonic scale shifts described here for the treble clef also apply to the bass clef for producing its notes as well. To get the next three measures of the treble clef, the diatonic scale shift S0 is applied to the last measure above (the third measure) to get the fourth measure, and then S1 is applied to the fourth measure to obtain the fth measure, and S0 is applied to the fth measure to get the sixth measure. This produces the fourth through sixth measures:

The nal three measures (measures 7 through 9) have an interesting interplay between the short and long timescale diatonic scale shifts. Measure 7 is obtained from measure 6 by applying the diatonic scale shift S1 :
Measure 6 Measure 7

S1

And then comes an interesting variation. The short timescale diatonic scale shift S0 is applied to the 3-note motive that ends the treble clef for measure 7, and then S1 is applied to the resulting 3-note motive, and then S1 is applied again. That produces the treble clef for measure 8. See Figure 1.17. As if to emphasize this variation, the half-notes in the bass clef, which in previous measures have been the only notes there, are converted into stacked quarter notes which creates 3-note chords at the beginning of each of the motives in measure 8.
Mathematics and Music, by James S. Walker.

1.4 Diatonic Transformations Scale Shifts


Measure 7
S0

25

Measure 8 E
S1 E S1 E

FIGURE 1.17. Diatonic scale shifts used to produce treble clef notes in Measure 8 of If I Aint Got You.

We conclude our analysis with measure 9. The treble clef of measure 9 is created by applying S1 to the last of the three note motives in measure 8 and then applying S0 twice:
Measure 8
S1

Measure 9 E
S0

S0

Remark 1.4.1. We cannot help but notice the way in which the number 3 (and also, to a degree 2) keeps reappearing in the melodic structure of Keys song. Notice that the initial melodic motive shown in (1.17) is composed of 3 notes. In fact, there is a micro-timescale of using diatonic scale shifts here: an application of S3 on the initial note, followed by an application of S2 on this second note. So, all together, there are in fact 3 timescale levels, and this combination of 3 timescales continues throughout each measure of the score. The short timescale repetitions of S0 are done 2 times in each measure to create 3 repetitions of a 3 note motive within each measure. The pattern of long timescale diatonic scale shifts from measures 1 to 7 is S0 S1 S0 S1 S0 S1

so S0 is used 3 times and S1 is used 3 times. In measure 8, where a variation is introduced, the diatonic scale shift S1 is used 2 times to create 3 repetitions of 3 note motives (which themselves begin with 3 note chords). Finally, 3 measures were created by using short timescale diatonic scale shifts (measures 1, 8, 9), while 3 measures were created by applying the long timescale diatonic scale shift S0 to a previous measure (measures 2, 4, 6), and 3 measures were created by applying the long timescale diatonic scale shift S1 to a previous measure (measures 3, 5, 7). The special number 3 could represent the fact that two together makes a third (their unity), which is appropriate for a love song. Moreover, the genre of this music (R & B) is inuenced by Gospel music, where 3 is of course of very special signicance (alluding to the Christian notion of the Trinity).

1.4.1

Compositions of Diatonic scale shifts

As shown in the preceding examples, diatonic scale shifts can be composed. That is, we can follow one diatonic scale shift by another one. This is the tonal sequencing that we referred to earlier. For example, suppose we just have one note, say C in the key of C-major. We could apply the diatonic scale shifts S1 followed by S2 to this note: C 1 D 2 F.
Mathematics and Music, by James S. Walker. S S

26

1. Algebra in Music

If we apply them in the opposite order, the end result is the same, the note F: C 2 E 1 F. We can write these calculations using parentheses: S1 (C) = D S2 (C) = E and and S2 (D) = F S1 (E) = F so so S2 S1 (C) = F S1 S2 (C) = F.
S S

We shall express S2 S1 (C) as S2 S1 (C), and S1 S2 (C) as S1 S2 (C). The symbol denotes composition of the diatonic scale shifts. We have found that S2 S1 (C) = F and S1 S2 (C) = F so we can write S2 S1 (C) = S1 S2 (C). In fact, because F = S3 (C), we have that S2 S1 (C) = S3 (C) and S1 S2 (C) = S3 (C). The reason, of course, is that these diatonic scale shifts are just invoking an elaborate form of the addition fact 1 + 2 = 2 + 1. This last statement implies that S2 S1 any sequence of notes = S1 S2 any sequence of notes because both sides of the equation are equal to S3 any sequence of notes . Equation (1.18) is usually expressed more succinctly as S2 S1 = S1 S2 . This equation is always understood to be an abbreviation of Equation (1.18). The discussion above illustrates the idea behind the proof of the following Theorem. Theorem 1.4.1 (Commutativity of Diatonic scale shifts). For all diatonic scale shifts, Sj and Sk , we have Sj Sk = Sk Sj (1.19) (1.18)

and both compositions are equal to Sj+k .


Remark 1.4.2. The diatonic scale shifts we have described are closely related to standard musical terminology. For example, the diatonic scale shift S2 is usually phrased as going up a 3rd , and the diatonic scale shift S4 is usually phrased as going up a 5th . In particular, S2 (C) = E and E is a 3rd up from C, while S4 (C) = G and G is a 5th up from C. Notice that S2 S2 = S4 . This situation is aptly described by the famous music theorist, David Lewin, as follows: The intervallic measurements . . . interact effectively with ordinary arithmetic. This obviates a defect in the traditional measurements which tell us, for example, that a 3rd and another 3rd compose to form a 5th . (3 + 3 = 5 ???).6

Exercises
1.4.1. For the set of three notes below:

David Lewin, Generalized Musical Intervals and Transformations, p. 17.


Mathematics and Music, by James S. Walker.

1.4 Diatonic Transformations Scale Shifts


perform the diatonic scale shift S1 . Write the result on this fragment of a staff:

27

1.4.2. For the set of three notes below:

perform the diatonic scale shift S2 . Write the result on this fragment of a staff:

1.4.3. Find at least ve diatonic scale shifts in the Mozart passage in Figure 1.18.

FIGURE 1.18. Introductory passage from 4th movement of Mozarts Divertimento No. 14, K.V. 270. 1.4.4. For the Beethoven passage shown in Figure 1.19, mark a sequence of diatonic scale shifts on the treble clef and on the bass clef. Within each sequence, identify a hierarchy of diatonic scale shifts. 1.4.5. Denote the set of diatonic scale shifts of notes by S, so S = {all diatonic scale shifts S0 , S1 , S1 , S2 , S2 , . . . }. Show that composition of diatonic scale shifts in S satises these four properties: (1) For each Sj , Sk in S, the composition Sj Sk is in S. (2) For each Sj , Sk , S in S, we have (Sj Sk ) S = Sj (Sk S ). (3) For each Sj in S, we have S0 Sj = Sj and Sj S0 = Sj . (4) For each Sj in S, there is a unique Sk in S that satises Sj Sk = S0 and Sk Sj = S0 .
Mathematics and Music, by James S. Walker.

28

1. Algebra in Music

FIGURE 1.19. Passage from Beethovens Piano Sonata No. 22 in F-Minor, Op. 54. Remark 1.4.3. Exercise 1.4.5 shows that the set of diatonic scale shifts S on notes is a mathematical group (c.f. Exercise 1.1.8 and Remark 1.1.3). Unlike the mathematical group H, however, the mathematical group S is innite. At least, it is potentially innite. The human ear only hears pitches that are within a certain range, and Sk would produce notes beyond this range when k is very large (k tending toward ) or very small (k tending toward ). But, at least in theory, we can view S as having innitely many members. The next exercise describes one way of resolving this issue. 1.4.6. In some areas of musical theory, only pitch classes are considered. In that case, two notes that differ by an octave are regarded as equivalent (in the same pitch class). Show that, if only pitch classes are considered, then the set of all diatonic scale shifts S can be viewed as just a nite set: S = {S0 , S1 , S2 , S3 , S4 , S5 , S6 }, and that the composition of diatonic scale shifts in S satises the properties: (1) For each Sj , Sk in S , the composition Sj Sk is in S . (2) For each Sj , Sk , S in S , we have (Sj Sk ) S = Sj (Sk S ). (3) For each Sj in S , we have S0 Sj = Sj and Sj S0 = Sj . (4) For each Sj in S , there is a unique Sk in S that satises Sj Sk = S0 and Sk Sj = S0 . so that S is a mathematical group (a nite one). Remark 1.4.4. This reduction to viewing only pitch classes is not universal in musical theory. On the one hand, if Mozart or Beethoven wrote notes beyond an octave it is reasonable to assume that they meant those notes and not their pitch class equivalents within a single octave. On the other hand, if we look at the relation between pitch and frequency in musical instruments, then pitches that are separated by an octave are certainly not equivalent. For one thing, they will differ in timbre.7 These two aspects, the use of pitches by composers and the physical/psychological aspects of how we hear tones, are no doubt related.

1.5

Diatonic Transformations Inversions, Palindromes

The other two types of diatonic scale transformations are inversions and palindromes. We will now describe examples of their use in music.

1.5.1

Diatonic Scale Inversions

When a set of notes is reected about a xed pitch on a scale, then we have a diatonic scale inversion.8 For example, consider the C-major scale: C, D, E, F, G, A, B, C. If we reect about the pitch at F (as if a horizontal mirror lies at the position for the note F), then a single pitch of A is 2 degrees above F on the scale,
Timbre is the psychological effect by which we distinguish the sound of different instruments. Although not completely understood, there is general agreement that timbre results from differing volume levels of pitch and overtone harmonics and their time variance within the attack and decay of notes. Our discussions of spectrograms in the next two chapters should further explain these ideas. 8 The term inversion is used frequently in music for many different purposes. It is also used in mathematics. So, to avoid confusion, we will always refer to the inversion described here by the somewhat cumbersome phrase, diatonic scale inversion.
Mathematics and Music, by James S. Walker.
7

1.5 Diatonic Transformations Inversions, Palindromes

29

and is reected to 2 degrees below, which is the note D. Or, the pitch at E is 1 degree below F, and is reected to +1 degree above, which is the note G. We shall denote this diatonic scale inversion as IF . Example 1.5.1 (J.S. Bachs Inventio 11). In Figure 1.20, we have indicated a simple diatonic scale inversion on a set of four notes from a passage by J.S. Bach. Here we have denoted the diatonic scale inversion by IE
I E

FIGURE 1.20. Measures 13 and 14 of Bachs Inventio 11, BWV 782, illustrating a diatonic scale inversion.

because the pitch used for reection is E . Notice that there is a marked on the treble clef for the position of the note E, indicating that notes marked at that position are E rather than E. The pitch E is for the third note in the sequence of 4 notes that we start with. These four notes are C D E C.

In terms of their degree changes on the scale away from E they are represented numerically as 2 1 0 2

which says that C is 2 notes below E , and D is 1 note below E , and E is 0 notes below E , and C is 2 notes below E . When IE is applied, the signs of these numbers are reversed, giving the sequence 2 1 0 2 which produces the notes G F E G as indicated on the score shown in Figure 1.20 (within the second box denoting the result of applying IE ). Reversing the signs of degree changes on a scale provides a way of detecting diatonic scale inversion without having to identify a mirror pitch. Consider this Bach example again. If we now write down the degree changes on the scale starting from the rst note C in our initial set of notes C then we obtain these three numbers +1 +1 2. That is, D is 1 degree on the scale above C, and E is 1 degree on the scale above D, and C is 2 degrees on the scale below E . Now, by reversing the signs, we have 1 and we see that the nal set of notes G F E G 1 +2 (1.20) D E C

Mathematics and Music, by James S. Walker.

30

1. Algebra in Music

has exactly these degree changes on the scale, starting from the initial note G. That is, F is 1 degree below G on the scale, E is 1 degree below F on the scale, and G is 2 degrees above E on the scale. The point of this second way of identifying an inversion is that we can start with any note and use the degree changes in (1.20) to create a sequence of notes that is a diatonic scale inversion of C D E C. For instance, if we start from the note D, then we have (the top line indicating degree changes on the scale): 1 D C 1 B +2 D

so D C B D is a diatonic scale inversion of C D E C.

1.5.2

Compositions of Diatonic Scale Inversions

Just like diatonic scale shifts, we can compose diatonic scale inversions. For instance, consider the following set of notes on the C-major scale (with the degree changes on the scale from one note to the next): 1 A G +3 C 3 G

Applying the diatonic scale inversion IB to these notes, we obtain since IB (A) = C :

+1 C D

3 A

+3 D

and then applying IC produces since IC (C) = C :

1 C B

+3 E

3 B

This calculation shows that IC IB A G C G = C B E B.

Now comes the really interesting part. If we compose these diatonic scale inversions in the opposite order we obtain: IB IC A G C G = F E A E so we conclude that IC IB A G C G = IB IC A G C G . This example shows that IC IB cannot be the same transformation as IB IC , which we abbreviate by writing IC IB = IB IC .
Mathematics and Music, by James S. Walker.

1.5 Diatonic Transformations Inversions, Palindromes

31

It is also worth observing from this work that composing two diatonic scale inversions has produced the same effect as a diatonic scale shift. For instance, we have IC IB A G C G since S2 A G C G We also have IB IC A G C G = S2 A G C G since they both produce the same note sequence, F E A E. Our discussion has shown the ideas behind the proof of the following Theorem. Theorem 1.5.1 (Non-Commutativity of Diatonic Scale Inversions). Let I and I denote diatonic scale inversions. If I = I , then I I = I I. = C B E B. = S2 A G C G

In fact, compositions of diatonic scale inversions are diatonic scale shifts: We have II = Sk and II = Sk for some integer k .
Before we leave the subject of diatonic scale inversions, we give another musical example, which shows that these transformations appear in more popular music as well. In this case, there is a diatonic scale inversion that plays an important part in a well-known traditional melody. Example 1.5.2 (Down in the Valley). In Figure 1.21 we show a diatonic scale inversion, I, acting on four notes in the melody of the traditional song, Down in the Valley. It is interesting to observe how this diatonic scale inversion emphasizes the corresponding lyrics of the song. The four notes that I is applied to correspond to the lyrics: Valley so low. While the resulting four notes after applying I correspond to the lyrics: Hear the wind blow. The musical contrast produced by the diatonic scale inversion emphasizes a contrast in the thoughts expressed by the lyrics. We leave it as an exercise for the reader to show that this diatonic scale inversion I satises the following equation: I = S3 IE . It is a composition of the diatonic scale shift S3 with the diatonic scale inversion IE .
I

Down in the val-

ley

Valley so low,

Hang your head o-

ver,

Hear the wind blow.

FIGURE 1.21. Diatonic scale inversion I in the melody of the traditional song, Down in the Valley.

1.5.3

Palindromes

Our nal type of scale transformation is palindrome symmetry. Like diatonic scale inversion, a palindrome symmetry can be thought of as using a mirror to reect notes. With a palindrome symmetry, however, the
Mathematics and Music, by James S. Walker.

32

1. Algebra in Music

FIGURE 1.22. A palindrome symmetry, P, of three notes produces a symmetrical motif of ve notes. The mirror line is marked by m.

mirror is placed vertically across the scale. The notes on the left side of the mirror are reected over to the right side of the mirror (if a note is on the mirror then it stays put). See Figure 1.22. Here are a couple of examples of palindrome symmetry in classical music. First, from a Chopin passage and then from a Bach piece. Example 1.5.3 (Chopins Valse, Op. 63 No. 3). In Figure 1.23, we show a palindrome symmetry on the bottom right of the score for a passage from Chopins Valse, Op. 63 No. 3. We have also indicated a hierarchy of diatonic scale shifts (some of which are modied diatonic scale shifts).
S

FIGURE 1.23. Measures 8 to 14 of Chopins Valse, Op. 63 No. 3, illustrating a palindrome symmetry, labeled by P with its mirror labeled m. (There is also a hierarchy of diatonic scale shifts labeled by S.)

Example 1.5.4 (J.S. Bachs Prelude and Fugue in E -major). J.S. Bach delighted in using all of the transformations we have described, including palindrome symmetries. In Figure 1.24 we show a passage of his that contains several palindrome symmetries. It is also notable in that it contains a concatenation of palindrome symmetries, a chain of palindrome symmetries. Besides occurring in classical music, palindromes also appear in more popular music. A famous example is a short bass run by Baghiti Khumalo within the song, You can call me Al by Paul Simon. Example 1.5.5 (Paul Simons You can call me Al). Within Paul Simons song, You can call me Al, there is a short sequence of bass notes that has a palindrome symmetry. We will show how this palindrome symmetry appears as a sequence of pitches and overtones within the music, using a graphical method that we will explain in detail in the next chapter. In Figure 1.25 we show a graphical depiction of a sound recording of the bass run, as played by the African bassist, Baghiti Khumalo. This display is called a spectrogram. We will describe it carefully in the next chapter. For now, it is enough to say that it provides a portrait of the volumes of the pitches and the overtones of the notes in the passage. The nearly perfect symmetry of the image about a vertical mirror line stands out clearly in the gure. Although the symmetry of the gure is not perfect, it is close enough to perfect for us to see that the notes in the score must have a perfect palindrome symmetry.
Mathematics and Music, by James S. Walker.

1.5 Diatonic Transformations Inversions, Palindromes

33

P1

CP

P2

FIGURE 1.24. Measures 78 to 83 of J.S. Bachs Prelude and Fugue in E major, illustrating two palindrome symmetries, labeled P1 and P2. The mirror for each palindrome is labeled by m. There is even a concatenation of palindromes, labeled CP and marked with a succession of mirror lines.

FIGURE 1.25. Palindrome symmetry within the song, You can call me Al. The display indicates the volumes of the note pitches, and their overtones, within a passage from the song. Brighter colors indicate higher volume. Pitch levels are marked along the vertical, with higher pitches appearing higher up than lower pitches, and time is marked along the horizontal. The notes that are played form a palindrome symmetry, with the mirror line marked by the line labeled m. It might be noticed that the display is very nearly, but not perfectly symmetrical. This is not a defect of the musicians technique. The notes of the score have a perfect palindrome symmetry, not the precise, machine recorded sound of the notes. V IDEO D EMO : Go to Videos link at book website and select 1.5.5: Palindrome in You Can Call Me Al.

Remark 1.5.1. This last example drew a distinction between notes on a score and the sound produced by playing those notes. In the next chapter we will discuss the method we have just shown of graphically analyzing
Mathematics and Music, by James S. Walker.

34

1. Algebra in Music

the sound of recorded music, nding that it closely corresponds to what we hear when music is played.

Exercises
1.5.1. Perform the diatonic scale inversion IB on this set of three notes:

1.5.2. Create a palindrome symmetry, positioning the mirror at the third note, for this set of three notes:

1.5.3. (a) Perform the the diatonic scale shift S2 on these three notes:

(b) Apply the diatonic scale inversion IB to the notes you found for part (a). (c) Now, on the original set of notes:

perform the diatonic scale inversion IB rst, followed by the diatonic scale shift S2 . (d) Explain why IB S2 = S2 IB . 1.5.4. In Figure 1.26, we show the beginning of the rst variation from Mozarts Twelve Variations on Ah vous dirais-je, Maman (also known as Twinkle Twinkle, Little Star), K.V. 265. Mark all of the diatonic scale shifts and diatonic scale inversions that you nd there (please ignore accidental markings on notes). 1.5.5. In Figure 1.27, we show the beginning of the eleventh variation from Mozarts Twelve Variations on Ah vous dirais-je, Maman, K.V. 265. Mark three palindromes on the treble clef for this portion of the score. 1.5.6. For the Mozart passage in Figure 1.27, nd an example of a hierarchy of diatonic scale shifts (diatonic scale shifts within diatonic scale shifts). Hint: It occurs on the the treble clef at the end. Retrograde transformation The retrograde transformation, R, is closely related to the palindrome symmetry, P. A retrograde transformation R applied to a sequence of notes will produce them in reverse order. For example, here is a retrograde transformation: R E

where we have shown R(G B E D) = D E B G. The following ve exercises deal with retrograde transformations.
Mathematics and Music, by James S. Walker.

1.5 Diatonic Transformations Inversions, Palindromes

35

FIGURE 1.26. Portion of a Mozart variation on Twinkle Twinkle, Little Star.

FIGURE 1.27. Portion of another Mozart variation on Twinkle Twinkle, Little Star. 1.5.7. Perform the retrograde transformation R on this set of three notes:

1.5.8. Find R(F B C D E). 1.5.9. Prove that R R = S0 .


Mathematics and Music, by James S. Walker.

36

1. Algebra in Music

1.5.10. Prove that R Sk = Sk R for all diatonic scale shifts Sk , and also that R I = I R for every diatonic scale inversion I. 1.5.11. One reason that retrograde transformations are used infrequently in musical compositions is that it is very difcult to perceive any connection between the sound pattern of the original set of notes and their sound pattern after retrograde transformation. For example, play this sequence of notes:9

This melody, if you can call it that, is the result of a retrograde transformation of a well-known melody. Which melody? If you cant recognize the tune, then that illustrates the point of this exercise.

1.6

Chromatic Transformations

In this section we discuss two mathematical transformations of notes on the chromatic scale. These transformations are analogous to the diatonic scales shifts and diatonic scale inversions discussed in the previous section, but they are done on the chromatic scale rather than on a diatonic scale.

1.6.1

Transpositions

A transposition is a uniform shift in pitch of a set of notes. It preserves the pitch relationships (semitone differences) between the notes of the melody. Often a transposition is done in order to play the notes in a different key. Because the pitch relationships are preserved, a transposed melody sounds equivalent to the original, just at a different overall pitch (a different key). Transposition is done by adding a xed whole number to each of the hour numbers for the notes on the Chromatic Clock. This shifts each of the hours by the same number of semitones and thereby achieves a transposition. Example 1.6.1. Here is the melody for a portion of the traditional song, Down in the Valley:

The tune is in the key of C-major. We will transpose it to the key of A-major. Here are the notes and hours on the Chromatic Clock: G 7 C 0 D 2 E 4 C 0 E 4 D 2 C 0 D 2 G 7 B 11 D 2 F 5 D 2 B 11 C 0 D 2 C 0

Since the starting note (tonic) for C-major is C with hour 0, and the tonic for A-major is A with hour 9, we simply add 9 to all of the numbers and read off the notes on the Chromatic Clock (remembering to subtract 12
9

If you dont have an instrument, you could use the M USE S CORE software listed in (1.1).
Mathematics and Music, by James S. Walker.

1.6 Chromatic Transformations


when we pass the top of the clock): 4 E 9 A 11 B 1 C 9 A 1 C 11 B 9 A 11 B 4 E 8 G 11 B 2 D 11 B 8 G 9 A 11 B 9 A

37

Here is the transposed melody, as it appears in the key of C-major:

Here is the precise denition of a transposition. Denition 1.6.1. A transposition is performed by adding a xed whole number to hour numbers on the Chromatic Clock. If the whole number k is added, then the transposition is denoted by Tk . From this denition it follows that the transposition used in Example 1.6.1 is T9 . There are precisely 12 distinct transpositions. They are the transpositions {Tk } for k = 0, 1, 2, . . . , 11. Example 1.6.2. We transpose the following brief melody

to the key of B . Its notes and hours on the Chromatic Clock are D 2 E 4 F 6 D 2 D 2 E 4 F 6 D 2

The melody is written in the key of D, and D has hour 2 on the Chromatic Clock. The tonic (starting note) for the key of B is B , which has hour 10 on the Chromatic Clock. To change from hour 2 to hour 10 we add 8 hours. So we apply the transposition T8 . We get these hours and their notes (written as they would appear in the original key of D): 10 0 2 10 10 0 2 10 B C D B B C D B Here is the transposed melody, as it appears in the key of D-major:

This last scoring is simplied if the key signature for B is employed. Using that key signature, the melody is written as follows:

Mathematics and Music, by James S. Walker.

38

1. Algebra in Music

Geometric Interpretation of Transpositions There is a beautiful geometric interpretation of transposition in terms of the Chromatic Clock. If we mark a point at each hour position on the clock and connect those points, we get a regular 12-sided polygon inscribed within the circle for the Chromatic Clock. See Figure 1.28. We shall refer to this polygon as the Chromatic 12-gon. C B A : B C : D D

D : E

G : A G F : G F

FIGURE 1.28. The Chromatic 12-gon. A regular 12-sided polygon inscribed within the circle for the Chromatic Clock.

There is an equivalence between transpositions and rotations of this Chromatic 12-gon about its center. For example, the following results hold (we leave their verication to the reader as an exercise): (a) A transposition from the key of C-major to the key of D-major corresponds to a rotation of the Chromatic 12-gon by 60 clockwise about its center, and this exactly preserves the shape of the Chromatic 12-gon. It is said to be a rotational symmetry of the Chromatic 12-gon. (b) A transposition of D-major to the key of G-major corresponds to a rotation of the Chromatic 12-gon by 150 clockwise about its center, and this is another rotational symmetry of the Chromatic 12-gon. From these results it is not hard to infer that every transposition corresponds to a rotation of the Chromatic 12-gon by a multiple of 30 about its center, and that altogether there are 12 distinct rotational symmetries of the Chromatic 12-gon. We record these observations as the following Theorem. Theorem 1.6.1. Every transposition corresponds to a rotation of the Chromatic 12-gon by a multiple of 30 about its center. Altogether, there are 12 distinct rotational symmetries of the Chromatic 12-gon.

1.6.2

Chromatic Inversions

There are scale inversions for the 12-note chromatic scale that are analogous to the diatonic scale inversions for the diatonic scales. These inversions cannot be described in terms of the note positions on clefs, however, since those clefs are not adapted to the chromatic scale (they are adapted to diatonic scales). The method of chromatic inversion is based on a simple modication of the idea of transposition. To see how the modication works, lets consider a specic transposition. For example, T9 . For T9 , any given hour
Mathematics and Music, by James S. Walker.

1.6 Chromatic Transformations

39

j for a note on the Chromatic Clock is mapped by T9 to a new hour by adding 9 to j. Equivalently, we could add j to 9: T9 (j) = 9 + j. If we now modify this formula by subtracting j from 9, then that will give us the method for doing a particular chromatic inversion. We will denote this chromatic inversion as R9 . So we have R9 (j) = 9 j. For R9 , any given hour j for a note on the Chromatic Clock is mapped by R9 to a new hour by subtracting j from 9. Here is an example. Example 1.6.3. The melody shown in Example 1.6.2 is D E F D D E F D

To chromatically invert this melody using R9 we proceed as follows. First, we write down the sequence of notes to be transformed, and their hours on the Chromatic Clock: D 2 E 4 F 6 D 2 D 2 E 4 F 6 D 2

Second, we subtract each of the hour numbers from 9, and then write down the notes for those hours: 7 G 5 F 3 D 7 G 7 G 5 F 3 D 7 G

Here is a scoring of this chromatically inverted melody (in the same key as the original):

Notice that the chromatic inversion has introduced notes that are not in the original key. We summarize our work in this example with the following denition. Denition 1.6.2. A chromatic inversion is performed by subtracting hour numbers on the Chromatic Clock from a xed whole number. If the hours are subtracted from the whole number k , then the chromatic inversion is denoted by Rk . There are precisely 12 distinct chromatic inversions. They are the chromatic inversions {Rk } for k = 0, 1, 2, . . . , 11. The same chromatic inversion can be notated in different ways. For example, the chromatic inversion R0 is cumbersome to use. It is the same as the chromatic inversion R12 , which is much easier to apply. Example 1.6.4. Suppose we apply the Chromatic Inversion R11 to the notes in the C-major chord, C-E-G. We have hours 0, 4, and 7, and we subtract them successively from 11, obtaining hours 11, 7, and 4. Those are the hours for the E-minor chord, E-G-B. So we have the mapping: (C-major chord) 11 (E-minor chord) As we state in the Theorem below, a chromatic inversion will always map major chords to minor chords, and map minor chords to major chords. On the other hand, transpositions always map major chords to major chords, and map minor chords to minor chords. For example, T5 maps the C-major chord C-E-G to the F-major chord F-A-C.
Mathematics and Music, by James S. Walker. R

40

1. Algebra in Music
Our discussion illustrates the ideas that lead to the following theorem.

Theorem 1.6.2. A transposition always maps major chords to major chords and minor chords to minor chords. A chromatic inversion always maps major chords to minor chords and minor chords to major chords. The set of transpositions and chromatic inversions acts transitively on the set of all major and minor chords in the following sense: For every pair of chords chord1 , chord2 , there is a transposition or chromatic inversion that maps chord1 to chord2 . The proof of this theorem is outlined in the exercises. Remark 1.6.1. Transpositions and chromatic inversions of standard 3-note chords can be diagrammed in an interesting way. For example, here are diagrams of a transposition T4 and chromatic inversion R11 of the C-major chord C-E-G: T4 G, 7 E, 4 C, 0
E 11, B E 8, G E 4, E

R11 G, 7
t  0  t  t  E, 4 t E 7, G   t t  t t 4, E C, 0 

11, B

This diagram is typical. Transpositions will map the notes of the chord, arranged in increasing hours around the Chromatic Clock, by keeping their same order. On the other hand, chromatic inversions will invert their order, in keeping with their description as inversions. Geometric Interpretation of Transpositions and Chromatic Inversions Previously we discussed how transpositions have a geometric interpretation as rotations of the Chromatic 12gon. We now show that there is a geometric interpretation of chromatic inversions in terms of mirror reection symmetries of the Chromatic 12-gon. Here is an example. Example 1.6.5. In Figure 1.29 we show a mirror drawn through the centers of opposite sides of the Chromatic 12-gon. If we reect through this mirror then notes are interchanged in pairs. For example, the notes A and C are switched. And the notes G and D are switched. We leave it as an exercise for the reader to show that this mirror reection produces the same effect on notes as the chromatic inversion R9 . Of course, this begs the question: Where did the 9 come from? The answer is that we look at where the note C with hour number 0 is reected to. It is reected to A with hour number 9. This example illustrates that each mirror symmetry of the Chromatic 12-gon corresponds to a chromatic inversion. There are 12 such mirror symmetries: 6 through opposite sides, and 6 through opposite corners, of the Chromatic 12-gon. Those 12 mirror symmetries match up precisely with the 12 chromatic inversions. We summarize our geometric interpretation of chromatic inversions, and also transpositions, with the following theorem. Theorem 1.6.3. The set of all transpositions and chromatic inversions corresponds exactly with the classical Euclidean symmetries of the 12-gon: the 12 distinct rotational symmetries of the 12-gon and the 12 distinct mirror reection symmetries of the 12-gon. The transpositions correspond to the rotational symmetries, and the chromatic inversions correspond to the mirror reection symmetries.
Mathematics and Music, by James S. Walker.

1.6 Chromatic Transformations

41

C B A : B C : D D

D : E

G : A G F : G F

FIGURE 1.29. Mirror on the Chromatic 12-gon for reecting notes in the Chromatic Scale, a chromatic inversion.

Exercises
1.6.1. Transpose the following three measures to the key of A-major (while remaining within the original key signature of C-major):

1.6.2. Transpose the following three measures to the key of B -major (while remaining within the original key signature):

1.6.3. Show that the transposition T7 converts the C-major chord C-E-G to the G-major chord G-B-D. 1.6.4. Explain why there are exactly 12 transpositions. 1.6.5. In Figure 1.30 we show the introduction to the Ray Charles song, Whatd I Say. There are several transpositions used in this passage. The rst transposition is indicated as T0 . Find the remaining transpositions that are indicated in the gure. (As explained in the caption for the gure, the transformation marked C is a combination of two separate transpositions.) 1.6.6 (Rotation of the Chromatic 12-gon). The following two facts were mentioned in the text, please solve them: (a) Show that a transposition from the key of C-major to the key of D-major corresponds to a rotation of the Chromatic 12-gon by 60 clockwise about its center, and that this exactly preserves the shape of the Chromatic 12-gon. It is said to be a rotational symmetry of the Chromatic 12-gon. (b) Show that a transposition of D-major to the key of G-major corresponds to a rotation of the Chromatic 12-gon by 150 clockwise about its center, and that this is another rotational symmetry of the Chromatic 12-gon.
Mathematics and Music, by James S. Walker.

42

1. Algebra in Music

E
T0

E
C

FIGURE 1.30. Opening measures of Ray Charles Whatd I Say, with some transpositions indicated. The transformation indicated by C is a combination of two separate transpositions. 1.6.7. Apply the chromatic inversion R4 to the following sequence of notes: C F G A

1.6.8. Apply the chromatic inversion R5 to the following sequence of notes: C F E G A B

1.6.9. Find the chord that results from applying the chromatic inversion R9 to the C-major chord, C-E-G. 1.6.10. Find the chord that results from applying the chromatic inversion R5 to the D-minor chord, D-F-A. 1.6.11. Find the chord that results from applying the transposition T9 to the C-major chord, C-E-G. 1.6.12. Find the chord that results from applying the transposition T5 to the D-minor chord, D-F-A. 1.6.13. Explain why there are exactly 12 chromatic inversions.
Mathematics and Music, by James S. Walker.

1.6 Chromatic Transformations


1.6.14. The following score fragment is taken from a composition by Mozart:

43

Find the chromatic inversion that occurs on the treble clef. 1.6.15 (Proof of Theorem 1.6.2). This exercise will outline the proof of Theorem 1.6.2. Work through the following 6 steps. (a) A major chord has intervals of 4 and 3 between its low, middle, and high notes. Show that every transposition applied to a chord will preserve those intervals. (That shows that transpositions always map major chords to major chords.) (b) Use the approach of item (a) to show that every transposition maps minor chords to minor chords. (c) A major chord has intervals of 4 and 3 between its low, middle, and high notes. Show that the chromatic inversion R0 applied to a chord will reverse those intervals so that the intervals are changed to 3 and 4 between the low, middle, and high notes, producing a minor chord. Then show, by similar reasoning, that R0 always maps a minor chord to a major chord. (d) Show that for each k = 0, 1, 2, . . . , 11, we have R0 Tk = Tk R0 then show that for each such k we have Rk = R0 Tk and Rk = Tk R0 .

(e) Combine the properties described in (c) and (d) to show that every chromatic inversion maps minor chords to major chords and maps major chords to minor chords. (f) Prove that the set of transpositions and chromatic inversions acts transitively on the set of major and minor chords. 1.6.16. Is there a geometric interpretation (using rotations and reections of a geometric gure) for diatonic scale shifts and diatonic scale inversions? If so, describe it. 1.6.17. The following is a portion of a score for a clarinet from Weberns Symphony Op. 21, Movement II:

Find all semitone changes, and express the second half of the passage as a composition of retrograde and transposition. Remark 1.6.2. This last passage by Webern is remarkable in that it is also a palindrome for the rhythm (note and rest durations), and the dynamics and articulation groupings for the playing of the passage. 1.6.18. Show that the set, P = {transpositions and chromatic inversions}, satises the following four properties of a mathematical group:

Mathematics and Music, by James S. Walker.

44

1. Algebra in Music

(1) For every pair of transformations P1 and P2 in P, the composition P1 P2 is also in P. (2) There is a transformation P0 in P that satises P0 P = P for every transformation P in P. (3) For all transformations P1 , P2 , and P3 in P, the following identity holds P1 P2 P3 = P1 P2 P3 . (4) For every transformation P in P, there is a transformation (denoted by P 1 ) that satises P 1 P = P0 and P P 1 = P0 . and P P0 = P

Show also that P is not a commutative group. That is, it is not always the case that P1 P2 equals P2 P1 for a given pair of transformations P1 and P2 in P. 1.6.19. Show that each transformation from the mathematical group P, dened in the previous exercise, that maps a given chord chord1 to a given chord chord2 is unique.

1.7

Web Resources

For additional study of basic music theory, there are a number of free sources available on the web. Here are four good ones: 1. A short introduction to music theory, with nice animation and basic sound examples, can be found at http://www.musictheory.net/ (1.21)

2. A fairly comprehensive discussion of music theory, with lots of great quotations from composers along with sound clips and scores, can be found at http://www.dolmetsch.com/introduction.htm 3. Free classical music scores can be found at this site: http://imslp.org/wiki/Main_Page (1.23) (1.22)

The page will display links near the top for browsing scores by composer, time period, or genre. It also has a search box on the left side. Some scores are protected by copyright, but many classical compositions are old enough that they are now public domain. 4. Some traditional songs and short classical pieces for piano can be found at this site: http://www.gmajormusictheory.org/Freebies/freebiesC.html (1.24)

Mathematics and Music, by James S. Walker.

Chapter 2

Trigonometry and Music

In order to fully understand many of the most important dimensions of musicsuch as pitch and timbre, as well as the nature and use of chordsit is necessary to study the mathematics of the trigonometric functions sine and cosine. These sinusoidal functions are a basic model for the frequency of pure tones, and sums of these sinusoidal functions are a basic model for the tones from all musical instruments. In this chapter we will explain the mathematical concepts of frequency and pitch and relate them to the musical concepts of overtones and chords. We will show how sums of sinusoidal functions can be used to model musical tones, and we also introduce the notion of localized sinusoidal analysis through the use of spectrograms. Spectrograms allow us to create visual representations of the time-frequency structure of music. After laying the foundations of analyzing music in a time-frequency manner in this chapter, we will apply time-frequency analysis to a wide variety of music in the next chapter.

2.1

Pitch and Frequency

There is a well-known connection between pitches in musical notes and the mathematical concept of frequency. In the 19th century, the famous German physicist, Helmholtz, did an experiment with tuning forks. He attached a pen to one of the tines of the fork and drew the fork across the paper while it was sounding a specic pitch. The vibration of the pen traced out a perfect sinusoidal wave. See Figure 2.1. The most fundamental aspect of a sinusoidal wave is that it repeats itself periodically. In physics, the distance from one peak of the wave to the next is called its wavelength. Another term for wavelength is cycle. We have marked one cycle for the sinusoid in Figure 3.1. The number of cycles that occur in 1 second is called the frequency. We have this formula for frequency: frequency = number of cycles . second

The unit of cycles/sec for measuring frequency is also called Hz, which is short for Hertz (another German physicist who did fundamental work in the study of frequency). For example, if the cycle shown in Figure 2.1 has a time duration of 0.025 seconds, then the frequency of the sinusoid is 1/0.025 = 40 Hz. Nowadays, with digital technology, we can record a representation of the sound wave from a tuning fork as a further demonstration of Helmholtzs idea. The sound wave from a tuning fork, or a musical instrument, creates oscillations of air pressure above and below the ambient air pressure. These oscillations cause the membrane of a microphone (or the human eardrum for that matter) to vibrate in synchrony. A transducer converts these membrane oscillations into an electrical current. Finally, electronic circuitry changes the electrical

46

2. Trigonometry and Music


1 cycle

FIGURE 2.1. Illustration of a famous experiment of Helmholtz. A pen is attached to one of the tines of a tuning fork. As the tuning fork is struck and drawn across a piece of paper at a uniform speed, the pen traces out a sinusoid. The distance marked between two peaks of the sinusoid is called a cycle. The number of cycles occurring in one second is the frequency of the sinusoid, and also of the pure tone produced by the tuning fork.

current into a digital le of integers. Typically these integers fall in the range between 2m and 2m 1 for some exponent m. The most common exponent is m = 15 used for 16-bit audio les (sometimes m = 7 is used for 8-bit les). In Figure 2.2 we show the plot of such a digital waveform recorded from a tuning fork. As described in the caption of Figure 2.2, the frequency for this pitch is 263 Hz. The note C in the 4th octave on the piano scale, denoted C4 , is known to have a frequency of 261.63 Hz. So the tuning fork is closely approximating the pitch for the note C4 . The approximation is so close (less then 2 Hz difference in frequency) that no one could hear the difference.1
8000 p 4000

4000

0.012

0.023

0.035

8000 0.046

FIGURE 2.2. Waveform from a recording of a tuning fork with one cycle marked, p = 0.0038 seconds. The frequency is about 1 cycle/0.0038 sec = 263 Hz, so the note of the tuning fork is C4 .

Just as the note C4 corresponds to the frequency 261.63 Hz, other notes have precise frequencies. For instance, the note A in the 3rd octave on the piano scale, A3 , has a pitch with frequency 220 Hz. In Table 2.1 we show correspondences between notes and frequencies for seven octaves of the piano scale (the 12-tone
It is a fascinating question how far apart in frequency pitches need to be in order for us to hear a difference. We will discuss this point later in the chapter.
Mathematics and Music, by James S. Walker.
1

2.1 Pitch and Frequency

47

equal-tempered scale). Table 2.1 is absolutely fundamental for our work in this chapter. It is important to
Table 2.1 P ITCH AND F REQUENCY C ORRESPONDENCES FOR E QUAL T EMPERED S CALE
Octave 1 C 33 C 35 D 37 D 39 E 41 F F

Frequency Ratio 1 21/12 2


2/12

Octave 2 C 65 C 69 D 73 D 78 E 82 F 87 F

Octave 3 C 131 C 139 D 147 D 156 E 165 F F

Octave 4 C 262 C 277 D 294 D 311 E 330 F 349 F

Octave 5 C 523 C 554 D 587 D 622 E 659 F F

Octave 6 C 1047 C 1109 D 1175 D 1245 E 1319 F F

Octave 7 C 2093 C 2218 D 2349 D 2489 E 2637 F 2794 F 2960 G 3136 G 3322 A 3520 A 3729 B 3951 C 4186

23/12 24/12 25/12 2 2


6/12 7/12

44 46 49

175 185 196

699 740

1397 1475

93 98

370 392

G 784 G 831 A 880 A 932 B 988 C 1047

G 1568 G 1661 A 1760 A 1865 B 1976 C 2093

28/12 29/12 210/12 211/12 2

G 52 A 55 A 58 B 62 C 65

G 104 A 110 A 117 B 124 C 131

G 208 A 220 A 233 B 247 C 262

G 415 A 440 A 466 B 494 C 523

Frequencies for pitches are rounded to nearest Hz, except for the A notes which are exact.

study it carefully.

2.1.1

Standardized frequency ratios

The rst column in Table 2.1 is included to emphasize the fact that the notes in each octave are all generated by the same frequency ratios, relative to the initial note C. In fact, if we let r stand for the number 21/12 = 1.059463 and C stand for the frequency of the C note that begins an octave, then the frequencies of each note in that octave have this pattern: C r0 C C r6 F C r12 C The nal note C, with frequency C r12 = C 2, is an octave above the initial note C, with frequency C . Notice how the powers of r correspond to hour positions on the Chromatic Clock. See Figure 2.3, and compare it with Figure 1.3. For example, the note A has frequency C r9 , and the power 9 is equal to the hour 9 position of the note A on the Chromatic Clock. For C r12 , however, we have a power of 12. The hour 12 on the Chromatic Clock is at the top, so to get the name of the note we would subtract 12 and get hour 0, giving the note C. We will write r12 r0 to indicate that these frequency multiples are giving the same notessame position on the Chromatic Clockbut the power r12 produces an octave higher pitched note than r0 . Lets go over some examples of relating powers of r to notes. Example 2.1.1 (Tuning note for the piano scale). The frequencies shown for the notes in Table 2.1 are based on A4 as the reference note. The frequency of 440 Hz is exact, as are the other frequencies for A notes. This
Mathematics and Music, by James S. Walker.

C r1 C C r7 G

C r2 D C r8 G

C r3 D C r9 A

C r4 E C r10 A

C r5 F C r11 B (2.1)

48

2. Trigonometry and Music

C B

r11
A : B

r12 r0

C : D

r1 r2 r3 r4
D

r10 r9 r8 r
G
7

D : E

G : A

r6
F : G

FIGURE 2.3. Frequency multipliers for notes arranged like the Chromatic Clock (c.f. Figure 1.3).

is not the case for the other notes, the frequencies shown for the other notes in the table are approximations to their exact values. For instance, the exact value for the frequency of C4 can be found from the fact that the frequency for A4 is C r9 , where C is the frequency for C4 . So we have C = (C r9 ) r9 1 = (440) 9 r 440 = 9/12 2 = 261.625565300599, and the frequency for C4 is 261.625565300599 Hz. The value of 262 Hz shown in Table 2.1 is just an approximation to the exact value, but so close (by far) that human ears could never detect any difference. Remark 2.1.1. It seems curious that the most basic scale is C-major (no sharps or ats), and yet the standard scale shown in Table 2.1 is tuned to the note A4 . Perhaps this has to do with orchestral tuning, since the pitch ranges of violins and trumpets, for instance, are more centered around 440 Hz than 262 Hz. Example 2.1.2 (Finding frequency of notes, using C as reference). As shown in the previous example, the note C4 has frequency 261.625565300599 Hz 262 Hz, and therefore C has frequency 4 261.625565300599 r1 = 261.625565300599 21/12 = 277.182630976873 277 Hz as shown in the table. And, G4 has frequency 261.625565300599 r7 = 261.625565300599 27/12 = 391.99543598175 392 Hz
Mathematics and Music, by James S. Walker.

2.1 Pitch and Frequency


also shown in the table.

49

2.1.2

Generating scales, transposing keys: Frequency multiplier method

The most important fact about Table 2.1 is that within each octave the ratios of frequencies for corresponding notes are the same. In other words, the frequency ratio in going up in pitch from, say, G to B, is the same for any octave. To see this, suppose we are in the 3rd octave. In that case, the frequency for C3 is C = 130.812782650299 Hz. The frequency ratio in going up in pitch from G3 to B3 is then C r11 (frequency for B3 ) = (frequency for G3 ) C r7 = r4 = 24/12 = 1.25992104989487. From this calculation, we can see that the value of C is irrelevant. We could have C be the value of the frequency for C in any of the octaves and we would always get r4 = 1.2599 . . . to be the frequency ratio in going from G to B in that octave. And it works the same for other pairs of notes. For hearing melodies, this result is quite important. As just one example, suppose two people, say a man and a woman, sing the same melody but an octave apart in pitch. The melody will sound the same, or at least equivalent, because our brains use the ratios of frequencies to measure the amount of change in pitch. Here are three applications of this ideawhich we call the frequency multiplier method to generating scales and transposing keys. Example 2.1.3 (Producing C-major scale). In Chapter 1, we described how to use addition on the Chromatic Circle to produce major scales. Here is how this is done using the frequency multiplier method. Suppose we want the C-major scale. We use the pattern of additions on the Chromatic Circle for major scales: 2 2 1 2 2 2 1

as powers of r and multiply frequencies (we have set C as a unit of 1 to save space): r0 C
r2

r2 D

r2

r4 E

r1

r5 F

r2

r7 G

r2

r9 A

r2

r11 B

r1

r12 r0 C

(2.2)

Example 2.1.4 (Changing Scales). In Chapter 1 we used addition on the Chromatic Circle to change from the scale of C-major to the scale of G-major. Here is how this is done using the frequency multiplier method. Since the frequency of G is r7 times the frequency of C, we simply multiply by r7 at the start of the calculations in (2.2), and remember that a frequency change by a factor of r12 gives r12 r0 (producing a note C): r7 G
r2

r9 A

r2

r11 B


r12 r0

r1

r0 C

r2

r2 D

r2

r4 E

r2

r6 F

r1

r7 G

(2.3)

Again, because our brains use the ratios of frequencies to measure the amount of change in pitch, the G-major scale in (2.3) will sound equivalent to the C-major scale in (2.2). The notes will sound different in every case, because they have different frequencies (different pitches). But the frequency ratios from note to note in each scale are the same, and our brains perceive this equivalence.
Mathematics and Music, by James S. Walker.

50

2. Trigonometry and Music

First three measures of Frere Jacques, key of C-major

Transposition to key of D-major

Transposition to key of A-major

FIGURE 2.4. First 3 measures of Frere Jacques. Top: In the key of C-major. Middle: Transposed to the key of D-major. Bottom: Transposed to the key of A-major.

Example 2.1.5 (Transposing between keys). A nice application of the frequency multiplier method is for performing key transpositions. In the top of Figure 2.4 we show the rst three measures of the traditional song, Frere Jacques, in the key of C-major. The notes of this brief melody are C D E C C D E C E F G.

To transpose this melody to the key of D-major, we rst write down its frequency multiples: r0 C r2 D r4 E r0 C r0 C r2 D r4 E r0 C r4 E r5 F r7 G (2.4)

Now, since D has a frequency multiple of r2 , relative to C, we multiply all of the powers of r in (2.4) by r2 and write down the notes corresponding to the new powers of r: r2 D r4 E r6 F r2 D r2 D r4 E r6 F r2 D r6 F r7 G r9 A (2.5)

The notes on the bottom line of (2.5) are the melody from Frere Jacques transposed to the key of D-major. In the middle of Figure 2.4 we show these notes on the treble clef. The melody sounds equivalent to the original, just in a different key. In fact, in this example, the notes in the two versions are so close in pitch that many listeners would not hear the difference (especially those who have not musically trained their hearing). Example 2.1.6 (Another key transposition). As another example of transposing between keys, lets transpose the melody from Frere Jacques to the key of A-major. The frequency multiples for the original melody are r0 C r2 D r4 E r0 C r0 C r2 D r4 E r0 C r4 E r5 F r7 G (2.6)

Since A has a frequency multiple of r9 , relative to C, we multiply all of the powers of r in (2.6) by r9 and write down the notes for these new powers of r: r9 A r11 B r13 r1 C r9 A r9 A r11 B r13 r1 C r9 A r13 r1 C r14 r2 D r16 r4 E (2.7)

Mathematics and Music, by James S. Walker.

2.1 Pitch and Frequency

51

The notes on the bottom line of (2.7) are the melody from Frere Jacques transposed to the key of A-major. On the bottom of Figure 2.4 we show these notes on the treble clef. As with the previous example, the melody sounds equivalent to the original, but in a different key. In this case, however, the notes in the A-major version are so different in pitch from the original that it is easy to hear the difference. We might say that the melody in A-major is in the wrong key. Perhaps it is more precise to say that the melody has been transposed to an equivalent melody in the key of A-major. Example 2.1.7 (Ray Charles Whatd I Say). A nice example of the use of transpositions in a musical passage occurs in the introduction to Ray Charles Whatd I Say. In Figure 2.5, we have indicated a sequence of transpositions occurring in the rst 12 measures of the Whatd I Say. The rst transposition is the identity transposition of multiplying by r0 = 1. This reproduces the notes in measures 1 and 2 to make up the notes in measures 3 and 4. Then by frequency multiplying by r5 , the notes in measures 3 and 4 are transposed up to make up the notes in measures 5 and 6. By multiplying by r5 , those notes are then transposed back down to the original set of notes in measures 7 and 8. Then an interesting splitting of the notes occurs. The rst ve notes in measure 7 are transposed up by multiplying by r7 , and the remaining six notes (in measure 7 and measure 8) are transposed up by multiplying by r5 . Those transpositions are then inverted back to the original set of notes by multiplying by r7 and r5 , which we have labelled as transformation C in Figure 2.5.

E
r0

E
r5 r 7

r 5

E
C

FIGURE 2.5. Opening measures of Charles Whatd I Say. The frequency multiplying operations are indicating transpositions of portions of the melody.

Remark 2.1.2 (Key Transpositions vis a vis Diatonic scale shifts). Looking at the three versions of the melody
Mathematics and Music, by James S. Walker.

52

2. Trigonometry and Music

from Frere Jacques in Figure 2.4, we see that each of the two transpositions to the keys of D-major and Amajor looks like one of the diatonic scale shifts discussed in Chapter 1 (S1 for the D-major case, and S5 for the A-major case). They are not diatonic scale shifts because the key signature was changed in each case. Neverthelessbecause addition in clock arithmetic underlies the Circle of Fifths and addition in clock arithmetic also corresponds to addition of powers in the frequency multiplier methodwe can write down the key transpositions on the treble clefs as if they are diatonic scale shifts, provided we also mark the key signatures that have changed as well. This is a beautiful example of the close connection between mathematics and music. One nal note on this topic. If we had done a diatonic scale shift of S1 on the Frere Jacques phrase while remaining in the key of C-major, then the notes would have been D E F D D E F D F G A.

producing a slightly different melody. This is an example of producing repetition with variation (the sequencing discussed in Chapter 1), rather than simply transposing to an equivalent melody. The examples we have given in this section, and the discussion in Remark 2.1.2, illustrate the ideas needed to prove the following theorem. Theorem 2.1.1. Addition on the Chromatic Clock is equivalent to multiplying frequencies of notes by powers of r = 21/12 .

Exercises
2.1.1. Use the frequency multiplier method to create the D-major scale. 2.1.2. Use the frequency multiplier method to create the A-major scale. 2.1.3 (Minor scale). Use the frequency multiplier method to create the natural B-minor scale. 2.1.4. Use the frequency multiplier method to create the natural F-minor scale. 2.1.5 (Frequencies in chords). In this gure we show the fundamental triadic (3-note) chords in the key of C-major:

I C

ii D

iii E

IV F

V G

vi A

vii B

I C

Using Table 2.1, nd the frequencies and frequency multiples (powers of r) for each of these chords. 2.1.6. Transpose the melody from Frere Jacques, shown at the top of Figure 2.4, to the key of F-major. 2.1.7. Transpose the melody from Frere Jacques, shown at the top of Figure 2.4, to the key of E-major. 2.1.8. Here is a short passage, on the treble clef, from Beethovens F r Elise: u

in the key of C-major. Transpose this passage to the key of F-major.


Mathematics and Music, by James S. Walker.

2.2 Instrumental Notes

53

2.1.9. Write out the chords that correspond to each of the boxed melodic motives that are transposed in the Whatd I Say score shown in Figure 2.5. 2.1.10. Which of the transpositions in Figure 2.5 are also diatonic scale shifts, and which are modied diatonic scale shifts? 2.1.11. On the Chromatic Clock, an increase of 7 hours corresponds to going up a fth. For example, the increase from hour 0 to hour 7 is going up from C to the fth note G in the key of C-major. Because of this, we say that multiplying by r7 is transposing up a fth. How would you describe the frequency multiplications by r7 , r5 , and r5 in the Whatd I Say score in Figure 2.5? How would you describe a frequency multiplication by r4 , and by r4 ? What frequency multiplications would correspond to transposing up a sixth and to transposing down a sixth?

2.2

Instrumental Notes

The notes from musical instruments, such as the human voice or a piano, violin, clarinet, or trombone, are much more complex than the notes from tuning forks. See the left sides of Figures 2.6 and 2.7 for examples of portions of waveforms from a ute and piano playing the note E4 . These waveforms have a fundamental period, at least approximately. We show this approximate fundamental period on the left side of Figure 2.6 for the ute note E4 , where it is easier to see in the graph. This fundamental period determines the frequency of approximately 329 Hz for the note E4 . We can see that these waveforms are not cycling in nearly so uniform a manner as the tuning fork waveform shown in Figure 2.2. In fact, as we shall examine more closely later in the chapter, they are combinations of several pure tones of differing frequency and loudness. For the note E4 , the pure tones that are combined have frequencies that are positive integer multiples of 329 Hz (called harmonics): 329 Hz, 2 329 Hz, 3 329 Hz, 4 329 Hz, 5 329 Hz, 6 329 Hz, . . . (2.8)

16000 8000 0 8000 16000 0.012 0.023 0.035 0.046

90 60 30 0 30 2632

658

1316

1974

FIGURE 2.6. Left: Waveform of portion of a ute note. The time p for an approximate cycle is p = 0.00304 seconds. So the fundamental frequency is approximately 329 Hz, which corresponds to the note E4 of frequency 329.63 Hz. Right: Magnitudes of harmonics within the ute note E4 , the harmonics of 329 Hz, 2 329 Hz, and 3 329 Hz, are clearly marked by spikes. Later in the chapter we will describe how this plot of magnitudes of harmonics is obtained. (Note: The small spike near 0, at around 65 Hz, is due to electrical oscillation from alternating current in the recording apparatus, not to the high pitched sound of the ute, so we are ignoring it here.)

For the ute note, the rst two harmonics of 329 Hz and 2 329 = 658 Hz are the loudest, the third harmonic 3 329 = 987 Hz is much fainter, and the higher multiples of 329 Hz are even fainter. See the right side of Figure 2.6. In the graph shown there, the heights of the peaks at 329 Hz, 2 329 = 658 Hz,
Mathematics and Music, by James S. Walker.

54

2. Trigonometry and Music

3 329 = 987 Hz, reect how loud those pitches would be if those pitches were heard separately. They are not heard separately, however. It is their combining together in creating the instrumental note E4 that provides that note with a much richer sound than a simple tuning fork pitch for E4 . Similarly, the piano note is a combination of harmonics. For the piano note, however, the graph on the right of Figure 2.7 shows that the magnitudes of the harmonics in (2.8) are much more equally distributed in size. An interesting feature of the plot of frequency magnitudes for the piano note is that the magnitude for the harmonic 2 329 Hz is actually larger than the magnitude for 329 Hz. Nevertheless, we refer to 329 Hz as the fundamental for this piano note since it corresponds to the pitch E4 that the note is sounding at. Moreover, it corresponds to the fundamental period of the piano notes waveform, which is equal to the fundamental period of the ute notes waveform, and both the ute and the piano are playing the same note E4 . We shall emphasize this point again later, as it is a subtle one: The fundamental harmonic for a note is the frequency that determines the notes pitch and that may or may not be the loudest harmonic in the waveform of the note.
16000 8000 0 8000 16000 0.012 0.023 0.035 0.046 90 60 30 0 30 2632

658

1316

1974

FIGURE 2.7. Left: Waveform of portion of a piano note E4 . Right: Magnitudes of harmonics within the piano note, the harmonics are marked by spikes at multiples of approximately 329.63 Hz. The rst spike is at 329 Hz, the second spike at 2 329 Hz, the third spike at 3 329 Hz, up to the seventh spike at 7 329 Hz.

The graphs of magnitudes of harmonics of ute notes and piano notes shown on the right of Figures 2.6 and 2.7 were obtained by computer processing of recordings of these notes. We will explain later in the chapter how this processing is done. There is a musical way, however, to see that the notes from instruments contain overtone vibrations. A classic demonstration can be done with a piano. For example, to demonstrate that the note C4 contains an overtone for G4 , you nd the middle C key on the piano and hold down the key for the rst G to the right of it. While holding down this key for G4 , so that the string for G4 is free to vibrate, you quickly strike the C4 key. As the C4 sound decays, you hear the G4 sound coming from the G4 string. The overtone vibration for G4 contained within the total sound vibration coming from the C4 string is causing the G4 string to vibrate (sympathetic vibration). If you have access to a piano, you should denitely try this experiment. And it works with other notes as well.2 We summarize this discussion with a denition of these frequency multiples in instrumental notes. Denition 2.2.1. For an instrumental note that contains frequencies of the form: o , 2 o , 3 o , 4 o , 5 o , 6 o , . . .

2 For simplicity, we phrased our discussion here as if there is a single string for each note on the piano. Most notes on a piano, including C4 and G4 , are produced from three strings that are struck simultaneously in order to produce more volume. This extra detail, however, does not change the point we are trying to make about overtones.

Mathematics and Music, by James S. Walker.

2.3 Frequency Content of Chords

55

The smallest frequency, o , is called the fundamental. The other frequencies are called overtones. All of the frequencies are called harmonics. The rst harmonic is o , the second harmonic is 2 o , the third harmonic is 3 o , and so on.
Remark 2.2.1. The physical explanation for why notes from musical instruments contain multiple harmonics is beyond the scope of this book. See Chapter 3 of Fourier Analysis, by James S. Walker (Oxford University Press, 1988) for a discussion of why stringed instruments have multiple harmonics. For other instruments, consult the book Music, Physics, and Engineering, by Harry F. Olson (Dover, 1967).

Exercises
2.2.1. Find the rst 6 harmonics of the note G2 , and the notes they correspond to (using Table 2.1). 2.2.2. Find the rst 6 harmonics of the note C2 , and the notes they correspond to (using Table 2.1). 2.2.3. Find the rst 6 harmonics of the note E2 , and the notes they correspond to (using Table 2.1). 2.2.4. Find the rst 6 harmonics of the note F2 , and the notes they correspond to (using Table 2.1). 2.2.5. Suppose the note C3 is struck on a piano. It has a frequency of 131 Hz. To what notes do its 2nd through 7th harmonics correspond (based on their pitch/frequency correspondences in Table 2.1)? 2.2.6. Suppose the note G3 is struck on a piano. It has a frequency of 196 Hz. To what notes do its 2nd through 7th harmonics correspond (based on their pitch/frequency correspondences in Table 2.1)? 2.2.7. Consider again the graph of the waveform of the ute note, shown in Figure 2.6. Explain from this graph, that there is a kind of half-period (a time duration equal to half of the period p shown in the gure), and that this half-period corresponds to the 2nd harmonic of 2 329 Hz. (Hint: the half-period is not an actual period, but remember that our hearing responds only to intensity of sinusoidals. We do not hear the signs of waveform values.)

2.3

Frequency Content of Chords

Chords form an integral part of music, especially Western music. In the previous section we saw that each instrumental note has a set of harmonics associated with itthe fundamental that determines the pitch of the note, and the overtones that provide depth, or richness, to the sound of the note.3 The set of harmonics for each instrumental note provides the basis for understanding how chords function. Lets look at a couple examples in detail. Example 2.3.1 (C-major chord). Suppose we have the C-major chord: C-E-G. In Table 2.2 we show the harmonics, up to frequency 4192 Hz, for the C-major chord in the 4th octave. There are several things to observe from this table. First, we observe that the 3rd harmonic of C4 nearly matches the 2nd harmonic of G4 . The 3rd harmonic for C4 is 786 Hz, which is only 2 Hz more than the 2nd harmonic for G4 of 784 Hz. There is a synergy, or reinforcement, between these two overtones, which enhances our perception of the sound. This synergy relates to the notion of consonance, an important area of musical theory which is still actively researched today. From now on, we will refer to these near matches of harmonics as consonances. Here is a description of the consonances shown in Table 2.2:
3 As we shall see later, human voices often display harmonics as well. A common theory in music is that musical instruments amplify and extend the capabilities of the human voice. Perhaps our perception of the depth, or richness, of instrumental notes is based on their connection to our voicesalthough the beauty of bird calls and other natural sounds cannot be discounted either.

Mathematics and Music, by James S. Walker.

56

2. Trigonometry and Music

Table 2.2

P ITCHES OF H ARMONICS FOR C- MAJOR CHORD

C4 (C4 ): (C5 ): (G5 ): (C6 ): (E6 ): (G6 ): (B ): 6 (C7 ): (D7 ): (E7 ): (G7 ): 262 524 786 1048 1310 1572 1834 2096 2358 2620 2882 3144 3406 3668 (B7 ): (C8 ):

E4 (E4 ): (E5 ): (B5 ): (E6 ): (G ): 6 (B6 ): (D7 ): (E7 ): (F ): 7 (G ): 7 (B7 ): 330 660 990 1320 1650 1980 2310 2640 2970 3300 3630 3960

G4 (G4 ): (G5 ): (D6 ): (G6 ): (B6 ): (D7 ): 392 784 1176 1568 1960 2352 2744 (G7 ): (A7 ): (B7 ): 3136 3528 3920

3930 4192

Based on Table 2.1. Notes in parentheses are on chromatic scale, consonances underlined.

1. There are 8 sets of consonances in Table 2.2. These eight sets of consonances can be listed according to how they lie among the harmonics of the three notes in the chord, as follows (an asterisk denotes an harmonic that is not in consonance): (1): {G5 , , G5 } (5): {D7 , D7 , D7 } (2): {E6 , E6 , } (6): {E7 , E7 , } (3): {G6 , , G6 } (7): {G7 , , G7 } (4): {, B6 , B6 } (8): {B7 , B7 , B7 }.

2. The root note of the chord, C4 , has consonances with both of the other two notes of the chord, E4 and G4 . 3. There are several harmonics consonant with the notes B and D. 4. There are some harmonics that are not consonant with notes in the key of C-major, and some notes that are not consonant with any of the notes on the chromatic scale. Item 1 shows that there are many consonances in this chord. Item 2 illustrates that the base note C4 of the chord provides substantial consonance with the other notes in the chord by nearly matching their harmonics. This major chord has a lot more consonances than the minor chord that we will analyze next. Item 3 shows that there is also consonance with two other notes, B and D. If the note B4 is sounded, then we would create additional synergy. The 2nd , 3rd , and 4th harmonics of B4 would be in consonance with harmonics of C4 , E4 , G4 . In fact, the chord C4 -E4 -G4 -B4 is a standard chord in music. It is a C-major seventh chord. Similarly, if the note D5 is sounded, then the 2nd and 3rd harmonics of D5 would be in consonance with
Mathematics and Music, by James S. Walker.

2.3 Frequency Content of Chords

57

harmonics of C4 , E4 , and G4 . The chord C4 -E4 -G4 -B4 -D5 is another standard chord in music. It is a C-major ninth chord. It is interesting to note that the C-major ninth chord is a simultaneous sounding of the C-major chord (chord I in the C-major key) and the G-major chord (chord V in the C-major key). This is related to the standard chord progressions I V and V I. Item 4 relates to the fact that the chromatic scale is not the only scale in music, nor can it be considered complete. The standard chords are not perfect in the sense of having only having consonances with notes in a given key, or even on the entire chromatic scale. Remark 2.3.1. As mentioned in Chapter 1, it is very common in music around the world to have scales that begin and end with notes that are an octave apart and to have a note that corresponds to a fth note on our standard Western scales. The reason for the universality of octave notes is easy to see. If a note is an octave higher than another note, then its frequency is twice that of the lower pitch note. So every harmonic of the octave higher note is a harmonic of the lower pitch note. One explanation for the predominance of fth notes on musical scales is that such notes have frequencies that are either exactly, or very closely approximate to, 3/2 times the frequency of the initial note. For example, on the piano scale, the fth note in the C-major scale is G, and G has frequency r7 times the frequency of the initial note C. But r7 = 27/12 = 1.49830707687668 1.5. Because the frequency multiplier for a note G is essentially equal to 3/2, the even harmonics of G will be consonant with harmonics of C. For example, the 2nd harmonic of G will be consonant with the 3rd harmonic of C because 2 frequency of G = 2 (C r7 ) 2 (1.5 C ) = 3 C = 3 frequency of C . Likewise, the 4th harmonic of G will be consonant with the 6th harmonic of C, the 6th harmonic of G will be consonant with the 9th harmonic of C, and so on. Furthermore, the odd multiples of the fundamental for G, the odd harmonics, fall neatly halfway in between consecutive harmonics of C. The consonance of harmonics of two notes with frequency ratio 3/2, or very close to 3/2, is the most probable reason for such a preponderance of fth notes in musical scales around the world. In the exercises, we shall expand on this point about r7 being closely approximate to 3/2, by examining other powers of r that also closely approximate simple fractions. These close approximations to simple fractions provide the harmonious sound of the equal-tempered (piano) scale. As another example of how frequency information for notes aids us in understanding chords, lets consider the D-minor chord (the ii chord for the C-major key). This will provide an interesting contrast with the major chord that we just examined. Example 2.3.2 (D-minor chord). In Table 2.3 we show harmonics for the three notes of the D-minor chord in the 4th octave. There are at least three points to note about the harmonics shown in Table 2.3: 1. There are 4 sets of consonances shown in Table 2.3. These four sets of consonances can be listed according to how they lie among the harmonics of the three notes in the chord: (1): {A5 , , A5 } (2): {A6 , A6 , , A6 } (3): {E7 , , E7 } (4): {A7 , A7 , A7 }.

Mathematics and Music, by James S. Walker.

58

2. Trigonometry and Music

Table 2.3

P ITCHES OF H ARMONICS FOR D- MINOR CHORD

D4 (D4 ): (D5 ): (A5 ): (D6 ): (F ): 6 (A6 ): (D7 ): (E7 ): (F ): 7 (A7 ):

F4 294 588 882 (F4 ): (F5 ): (C6 ): (F6 ): (A6 ): (C7 ): (F7 ): (G7 ): (A7 ): (C8 ): 349 698 1047 1396 1745 2094 2443 2792 3141 3490 3839 4188

A4 (A4 ): (A5 ): (E6 ): (A6 ): (C ): 7 (E7 ): (A7 ): (B7 ): 440 880 1320 1760 2200 2640 3080 3520 3960

1176 1470 1764 2058 2352 2646 2940 3234 3528 3822

Based on Table 2.1. Notes in parentheses are on chromatic scale, consonances underlined.

2. In contrast to the C-major chord, there are no consonances with harmonics of the middle note, F4 , of the chord. 3. There are 3 harmonics for F4 that are consonant with the note C. Items 1 and 2 show that the base note D4 provides consonance for only the highest note A4 in the chord. This could be interpreted as a kind of weakness, or structural instability, of the chord, because there is no reinforcement of the middle notes harmonics by harmonics of the other notes. Furthermore, the consonance of harmonics of both D4 and F4 with only the highest note of the chord, A4 , could be an explanation for the plaintive sound of the chord (a property shared by other minor chords). Item 3 shows that there is also consonance with the note C. If the note C5 is sounded, then we would create additional synergy, due to consonance of the 2nd , 3rd , and 4th harmonics of C5 with harmonics of F4 . In fact, the chord D4 -F4 -A4 -C5 is a standard chord in music. It is a D-minor seventh chord. Notice also that the presence of the pitches E6 and E7 , within the harmonics of A4 and D4 , implies that there is consonance within the chord D4 -F4 -A4 -C5 -E5 . This chord is called a D-minor ninth chord.

Exercises
2.3.1. The F-major chord in the key of C-major is F-A-C. In Table 2.4 we show frequencies for the harmonics of the notes F3 , A3 , C4 , similar to what we did in Table 2.2. Within the parentheses in this table, write in the notes from the equal-tempered scale that correspond to the harmonics in Table 2.1, and underline the consonances. Provide an analysis of this chord as was done in Example 2.3.1. 2.3.2. The G-major chord in the key of C-major is G-B-D. In Table 2.5 we show frequencies for the harmonics of the notes G3 , B3 , D4 , similar to what we did in Table 2.2. Within the parentheses in this table, write in the notes from the equal-tempered scale that correspond to the harmonics in Table 2.1, and underline the consonances. Provide an analysis of this chord as was done in Example 2.3.1.
Mathematics and Music, by James S. Walker.

2.3 Frequency Content of Chords

59

Table 2.4
F3 ( ( ( ( ( ( ( ( ( ( ): ): ): ): ): ): ): ): ): ): 175 350 525 700 875 1050 1225 1400 1575 1750 1925 ( ): 2100 2275 2450 ( ( ): ): 2625 2800

H ARMONICS FOR F- MAJOR CHORD


A3 ( ( ( ( ( ( ( ( ( ( ): ): ): ): ): ): ): ): ): ): 220 440 660 880 1100 1320 1540 1760 1980 2200 2420 ( ): 2640 ( ( ( ( ( ( ( ( ( ( C4 ): ): ): ): ): ): ): ): ): ): 262 524 786 1048 1310 1572 1834 2096 2358 2620

Table 2.5
G3 ( ( ( ( ( ( ( ( ( ( ): ): ): ): ): ): ): ): ): ): 196 392 588 784 980 1176 1372 1568 1764 1960 2156 ( ): 2352 2548 2744 ( ( ): ): 2940 3136

H ARMONICS FOR G- MAJOR CHORD


B3 ( ( ( ( ( ( ( ( ( ( ): ): ): ): ): ): ): ): ): ): 247 494 741 988 1235 1482 1729 1976 2223 2470 2717 ( ): 2964 ( ( ( ( ( ( ( ( ( ( D4 ): ): ): ): ): ): ): ): ): ): 294 588 882 1176 1470 1764 2058 2352 2646 2940

Mathematics and Music, by James S. Walker.

60

2. Trigonometry and Music

2.3.3. Your results for the previous two exercises should correspond exactly (except for note names) with the results for Example 2.3.1. Explain mathematically why this is so. 2.3.4. Show that the D-minor chord in the key of C-major is the rst chord in the key of D-minor (by rst chord, we mean that it has D as its base note, the starting note of the D-minor key). 2.3.5. The E-minor chord in the key of C-major is E-G-B. In Table 2.6 we show frequencies for the harmonics of the notes E3 , G3 , B3 , similar to what we did in Table 2.3. Within the parentheses in this table, write in the notes from the equal-tempered scale that correspond to the harmonics in Table 2.1, and underline the consonances. Provide an analysis of this chord as was done in Example 2.3.2. Table 2.6
E3 ( ( ( ( ( ( ): ): ): ): ): ): 165 330 495 660 825 990 1155 ( ( ( ): ): ): 1320 1485 1650 1815 ( ): 1980 2145 ( ): ( ( ( ): ): ): ( ( ( ( ( (

H ARMONICS FOR E- MINOR CHORD


G3 ): ): ): ): ): ): 196 392 588 784 980 1176 1372 1568 1764 1960 2156 2352 ( ( ): ): ( ( ( ( ( ( B3 ): ): ): ): ): ): 247 494 741 988 1235 1482 1729 1976 2223

2.3.6. The A-minor chord in the key of C-major is A-C-E. In Table 2.7 we show frequencies for the harmonics of the notes A3 , C4 , E4 , similar to what we did in Table 2.3. Within the parentheses in this table, write in the notes from the equal-tempered scale that correspond to the harmonics in Table 2.1, and underline the consonances. Provide an analysis of this chord as was done in Example 2.3.2. 2.3.7. Your results for the previous two exercises should correspond exactly (except for note names) with the results for Example 2.3.2. Explain mathematically why this is so. 2.3.8. The diminished B-minor chord in the key of C-major is B-D-F. Produce a table like Table 2.3 for the notes B3 , D4 , F4 . Provide an analysis of this chord as done in Example 2.3.2. 2.3.9. An arpegiatted chord is a chord where the notes are played in close, generally overlapping, succession. Here is an example of an arpeggiated chord of three notes:

Would you say this chord is major or minor? Identify the notes (including octaves) for this chord.
Mathematics and Music, by James S. Walker.

2.3 Frequency Content of Chords

61

Table 2.7
A3 ( ( ( ( ( ( ): ): ): ): ): ): 220 440 660 880 1100 1540 1760 ( ( ( ): ): ): 1980 2200 2420 2640 ( ): 2860 3080

H ARMONICS FOR A- MINOR CHORD


C4 ( ( ( ( ( ( ): ): ): ): ): ): 262 524 786 1048 1310 1834 2096 ( ( ( ): ): ): 2358 2620 2882 3144 ( ): 3406 ( ( ): ): ( ( ( ( ( ( E4 ): ): ): ): ): ): 330 660 990 1320 1650 2310 2640 2970 3300

2.3.10. The following short motif

is from Ray Charles Whatd I Say. (a) The triadic chord in the middle of the motif is G2 -E4 -C5 . Compile a table of the frequencies of the harmonics of these notes, and do an analysis of this chord along the lines of the analyses in Examples 2.3.1 and 2.3.2. (b) Following this triadic chord, there is an octave chord of G2 -G4 . Which harmonics of G2 are amplied by playing the G4 note along with it? Explaining harmonies in the equal-tempered scale In Remark 2.3.1, we showed that r7 is closely approximate to 3/2 and that this explains the consonances of a fth note on a major scale with the starting note (for example, the consonances of G with C in the C-major scale). The following four exercises show that this occurs for other powers of r in the frequency multipliers used for the equal-tempered scale. 2.3.11. Show that r4 is closely approximate to 5/4. Explain how this provides consonance of harmonics of E4 with harmonics of C4 . And, in general, why the third note on a major scale will have consonances with harmonics of notes of the rst note on the major scale. 2.3.12. Show that r5 is closely approximate to 4/3. Explain how this provides consonance of harmonics of F4 with harmonics of C4 . And, in general, why the fourth note on a major scale will have consonances with harmonics of notes of the rst note on the major scale. 2.3.13. Based on the results of the previous two exercises, which note would you say is more harmonious with the starting note of a major scale, the third note or the fourth note?
Mathematics and Music, by James S. Walker.

62

2. Trigonometry and Music

2.3.14. Show that r9 is closely approximate to 5/3. Explain how this provides consonance of harmonics of A4 with harmonics of C4 . And, in general, why the sixth note on a major scale will have consonances with harmonics of notes of the rst note on the major scale. 2.3.15 (Power chords). Occasionally in music, a two-note chord is played, usually the notes in the chord will differ by a fth (as in the two-note chord C4 -G4 ). In rock music, where this is done more frequently than in other genres, such a two-note chord is called a power chord. Using Table 2.2, look at the consonances of the power chord C4 -G4 . What musical reasons can you think of for playing this chord? 2.3.16 (Pentatonic chords). In Exercise 1.3.4, we introduced the Pentatonic-C scale. The C-major chord is a chord that could be played with this scale. But what about other chords? Produce a table like Table 2.2 for the chord D4 -E4 -G4 and analyze the consonances for this chord. Do the same for the chord D4 -G4 -A4 . What conclusions can you draw from these analyses (especially about musical reasons why chords are not generally used with pentatonic scales)? 2.3.17 (Cents). Since frequency ratios in the equal tempered scale increase exponentially, as powers of r = 21/12 , it makes sense to measure the difference between any two frequencies in a logarithmic way. The cents measure of difference between two frequencies 1 and 2 is (cents) = 1200 log2 (2 /1 ) = 1200 log(2 /1 )/ log(2). (2.9)

(a) Show that the difference in cents between each pair of successive notes on the equal-tempered scale is 100. (b) Show that the difference in cents between the frequencies 3951 Hz (B7 on the equal-tempered scale) and 3930 Hz (the 15th harmonic of C4 in the rst column of Table 2.2) is 9.23 cents. Remark 2.3.2 (Discriminating Differences in Frequencies). The difference of 9 cents found in part (b) of the previous exercise is less than one-tenth of the difference in cents for frequencies between successive notes on the equal tempered scale. If two notes of fundamental frequencies, 3951 Hz and 3930 Hz were played separately it would be difcult for most listeners to hear any difference. If they are played as 15th harmonics of an instrumental note C4 , then they are essentially equivalent. This is why we listed B7 as a note corresponding to the overtone harmonic 3930 Hz in the rst column of Table 2.2. Similar remarks apply to other harmonics in Tables 2.2 and 2.3 which differ by only a few cents from the notes indicated for them on the equal tempered scale.4 Other scales Besides the twelve-tone equal tempered scale, there are many other scales. Chapter 5 of the book Music: A Mathematical Offering, by David Benson, Cambridge University Press, 2007, discusses several additional scales. Additional scales are examined in the book Tuning, Timbre, Spectrum, Scale, by William Sethares, Springer Publishing, 2005. Here we will just look at one different type of scale, the just intonation scale. See Table 2.8. Just intonation is used by many male singing groups (Barbershop quartets) and has been used by composers as well (such as Ben Johnston). In the following four exercises we examine the just intonation scale. 2.3.18. Using Table 2.9, analyze the chord C4 -E4 -G4 on the just intonation scale. How do your results compare with our analysis in Example 2.3.1? 2.3.19. Using Table 2.10, analyze the chord D4 -F4 -A4 on the just intonation scale. How do your results compare with our analysis in Example 2.3.2? 2.3.20 (Key transposition in just intonation). One difculty that is often cited with the just intonation scale is the problem of key transposition. Transpose the C-major scale in the 4th octave of the just intonation scale by multiplying all of its notes by the frequency ratio 3/2 (so that the transposed scale begins with G4 ). How does this transposed scale compare (using cents as a measure of frequency difference) with the frequencies shown in Table 2.8 for the notes that should be on a G-major scale beginning with G4 ? Which notes differ the most?
4 Further details on detecting differences of frequencies for pitches and harmonics can be found in Chapter 4 of the book Music: A Mathematical Offering, by David Benson, Cambridge University Press, 2007, and throughout the book Tuning, Timbre, Spectrum, Scale, by William Sethares, Springer Publishing, 2005.

Mathematics and Music, by James S. Walker.

2.3 Frequency Content of Chords

63

Table 2.8
Frequency Ratio 1 16/15 9/8 6/5 5/4 4/3 7/5 3/2 8/5 5/3 7/4 15/8 2

P ITCH AND F REQUENCY C ORRESPONDENCES FOR J UST S CALE


Octave 1 C 32 D 34 D 36 E E F F G 38 40 43 45 48 Octave 2 C 64 D 68 D 72 E E F F G 77 80 85 90 96 Octave 3 C 128 D 137 D 144 E E F F G 154 160 171 179 192 Octave 4 C 256 D 273 D 288 E E F F G 307 320 341 358 384 Octave 5 C 512 D 546 D 576 E E F F G 614 640 683 717 768 Octave 6 C 1024 D 1092 D 1152 E E F F G 1229 1280 1365 1434 1536 Octave 7 C 2048 D 2185 D 2304 E E F F G 2458 2560 2731 2867 3072

A 51 A 53 B 56 B 60 C 64

A 102 A 107 B 112 B 120 C 128

A 205 A 213 B 224 B 240 C 256

A 410 A 427 B 448 B 480 C 512

A 819 A 853 B 896 B 960 C 1024

A 1638 A 1707 B 1792 B 1920 C 2048

A 3277 A 3413 B 3584 B 3840 C 4096

Frequencies for C notes are exact. If needed, other frequencies are rounded to nearest Hz.

Table 2.9
C4 ( ( ( ( ( ( ( ( ( ( ): ): ): ): ): ): ): ): ): ): 256 512 768 1024 1280 1536 1792 2048 2304 2560 2816 ( ): 3072 3328 3584 ( ( ): ): 3840 4096

H ARMONICS FOR C- MAJOR CHORD , JUST SCALE


E4 ( ( ( ( ( ( ( ( ( ( ): ): ): ): ): ): ): ): ): ): 320 640 960 1280 1600 1920 2240 2560 2880 3200 3520 ( ): 3840 ( ( ( ): ): ): ( ( ( ( ( ( G4 ): ): ): ): ): ): 384 768 1152 1536 1920 2304 2688 3072 3456 3840

2.3.21 (Key transposition in just intonation). The previous exercise showed a small problem with transposition in just intonation. Notice that G-major is right next to C-major on the Circle of Fifths. There is a much greater problem in transposing from C-major to a key that is a lot farther away on the Circle of Fifths, such as B-major. Transpose the C-major scale in the 4th octave of the just intonation scale by multiplying all of its notes by the frequency ratio 15/8 (so that the transposed scale begins with B4 ). How does this transposed scale compare (using cents as a measure of frequency difference) with the frequencies shown in Table 2.8 for the notes that should be on a B-major scale beginning
Mathematics and Music, by James S. Walker.

64

2. Trigonometry and Music

Table 2.10
D4 ( ( ( ( ( ( ): ): ): ): ): ): 288 576 864 1152 1440 1728 2016 ( ( ( ): ): ): 2304 2592 2880 3168 ( ): 3456 3744

H ARMONICS FOR D- MINOR CHORD , JUST SCALE


F4 ( ( ( ( ( ( ): ): ): ): ): ): 341 682 1023 1364 1705 2046 2387 ( ( ( ): ): ): 2728 3069 3410 3751 ( ): 4092 ( ( ): ): ( ( ( ( ( ( A4 ): ): ): ): ): ): 427 854 1281 1708 2135 2562 2989 3416 3843

with B4 ? Which notes differ the most? Remark 2.3.3. What the previous two exercises reveal is that with just intonation it is impossible to transpose between keys, especially keys that are far apart on the Circle of Fifths, and preserve the pitch changes in melodies. Historically, the equal tempered scale was developed as a compromise solution to the problem of having good harmony while also allowing transposition between keys.

Mathematics and Music, by James S. Walker.

Appendix A

Solutions to Odd-Numbered Exercises

Chapter 2
2.1.1 2.1.3 2.1.5 E F F G C D G A B C C D A

D E F B C

A B E

F G

2.1.7 (a) We have these calculations on the Chromatic Clock: 9 A 11 B


+2


12 at top

+1

0 C

+2

2 D

+2

4 E

+1

5 F

+2

7 G

+2

9 A.

(b) Because the sequence of adds is the same as for C-major, but starting with A in (1.9). (c) E F G A B C D E (d) B C D E F G A B (e) The adds on the Chromatic Clock for a natural minor scale are +2 +1 +2 +2 +1 +2 +2

while these adds applied to the fth note produce +2 +2 +1 +2 +2 +1 +2

and, by comparing these two sequences of adds, we see that there is only one new note introduced with the second set of adds, and it introduces a new note that is two notes ahead of the new starting note (the fth note) on the Chromatic Clock. For example, two notes ahead of E on the Chromatic Clock is the note F as found in (c), and two notes ahead of B is C as we found in (c), and so on. 2.2.1 2.2.3 F A F A C E G D G B B F A E D G E F C A G A A C B G B B C E A E C D B

Treble Clef: Bass Clef:

G F

C D

E B D

2.2.5

A C A

A F C A

F F A

B D G E

66
2.2.7 Treble clef: G

APPENDIX A. SOLUTIONS TO ODD-NUMBERED EXERCISES


B D F E C G C B G F Alto clef: E Bass clef: A C A G E B D A C E F C F F E

Tenor clef: F G

E B D

2.2.9 The equations with fractions for the three measures are 1 1 1 + + + 8 8 4 1 + 2 3 1 1 + 16 32 1 1 + 8 16 + + 1 1 4 + = 16 4 4

1 1 1 1 4 + + + = 32 8 8 8 4

1 1 1 1 1 1 1 4 + + + + + + = 32 32 8 4 8 8 4 4

2.2.11 The equations with fractions for the three measures are 3 1 1 1 1 + + + = 8 8 4 4 4 1 1 1 1 3 1 1 + + + + + = 8 8 16 16 8 4 4 1 + 32 1 1 + 16 32 + 1 1 + 16 32 + 1 + 32 1 1 + 32 64 + 1 + 64 1 1 + 16 32 + 1 1 1 3 + + = 32 16 4 4

2.2.13 For the rst measure, the equation is 1 8 + 1 + 8 1 4 + 1 1 + 4 8 + 1 4 = . 8 4

1 1 For the rst triplet, each note is 24 duration. For the second triplet, each note is 12 duration. For the second measure, the equation is 4 (1) = . 4 1 Each note in the triplet is 3 duration. For the third measure, the equation is 4 (1) = . 4 1 The rst two notes in the triplet are 3 duration. The last two notes are each 1 duration. 6

2.2.15 For the rst measure, the equation is 1 2 + 1 2 = 4 . 4


1 6

1 For the quintuplet, the notes are each 10 duration. For the triplet, the rst two notes are each 1 two notes are each 12 duration. For the second measure, the equation is

duration, the second

1 2

1 + 4

1 8

1 4 = . 8 4
1 40

1 For the septuplet, the notes are each 14 duration. For the quintuplet, the notes are each For the third measure, the equation is 1 1 4 1 + + = . 2 4 4 4

duration.

1 1 For the quintuplet, the rst note is duration 10 , the second two notes are each 20 duration, and the last two notes are 1 1 1 each 10 duration. For the triplet, the rst note is 12 duration, and the second note is 6 duration.

Mathematics and Music, by James S. Walker.

67
2.2.17 For the Adagio tempo, a quarter note is 3 seconds. For the Allegro tempo, a quarter note is 2 Andante tempo, a quarter note is 3 seconds. 4 2.2.19 a. Letting x stand for the remaining time in the rst measure, we have 1 1 + 16 32 which yields x= 21 32 16 + 4 + 1 = 32 1 1 1 + + = 32 8 2 +x= 3 4
1 2

seconds. For the

so there is a 1/32nd rest, followed by a 1/8th rest, followed by a 1/2 rest. b. Letting x stand for the remaining time in the second measure, we have 1 + 16 which yields x= 5 16 4+1 = 16 1 1 = + 4 16 1 1 + 4 8 +x= 3 4

so there is a 1/16th rest, followed by a 1/4th rest. c. Letting x stand for the remaining time in the third measure, we have 1 1 + 16 32 which yields x= 29 32 16 + 8 + 4 + 1 = 32 1 1 1 1 = + + + 2 4 8 32 +x= 4 4

so there is a 1/32nd rest, followed by a 1/8th rest, followed by a 1/4th rest, followed by a 1/2 rest. 2.3.1 The 7 sharp key signature is

which is C -major, or A -minor. Its notes are C , D , E , F , G , A , B , C . All the notes are now sharps, so no further key signatures with sharps appear on the Circle of Fifths. The notes E and B are enharmonic with the notes F and C, respectively.
Mathematics and Music, by James S. Walker.

68

APPENDIX A. SOLUTIONS TO ODD-NUMBERED EXERCISES


The 7 at key signature is

which is C -major, or A -minor. Its notes are C , D , E , F , G , A , B , C . All the notes are now ats, so no further key signatures with ats appear on the Circle of Fifths. The notes C and F are enharmonic with the notes B and E, respectively. 2.3.3 (a) C-major (I): C-E-G D-minor (ii): D-F-A E-minor (iii): E-G-B F-major (IV): F-A-C G-major (V): G-B-D A-minor (vi): A-C-E B -minor (vii ): B-D-F (b) I, IV, and V all have intervals of 4 and 3 between notes, while ii, iii, and vi all have intervals of 3 and 4 between notes. The diminished minor vii has intervals of 3 and 3 between notes, which differs from the other minor chords (as well as major chords). (c) The G-major scale is G A B C D E F G so the iii chord starts on B with 11 B and that gives the B-minor chord, B-D-F . 2.3.5 The chord C-E-G is at hours 0, 4, and 7 on the Chromatic Clock, so its intervals are 4 and 3. Using the note D at hour 2, the chord with intervals 4 and 3 would be D-F -A, but F is not part of the pentatonic C-scale. Or, using the intervals 3 and 4, the chord would be D-F-A, but F is not part of the pentatonic C-scale. With the note E and intervals of 4 and 3, the chord would be E-G -B, but G is not on the pentatonic C-scale. Or, with intervals of 3 and 4, the chord would be E-G-B, but B in not on the pentatonic C-scale. 2.4.1 The solution is shown in the following gure:
12 at top +3

2 D

+4

6 F .

2.4.3 A solution is shown in Figure A.1. 2.4.5 (1) Since Sj Sk = Sk+j , and Sk+j is in S, we have Sj Sk in S. (2) We have (Sj Sk ) S = Sj+k S = S(j+k)+ = Sj+(k+) = Sj (Sk S ). (3) We have S0 Sj = S0+j = Sj and Sj S0 = Sj+0 = Sj . (4) Since Sj Sk = Sk+j , we must require that j + k = 0 in order to have Sj+k = S0 . But, j + k = 0 implies that j = k. Similarly, Sk Sj = S0 implies that j = k. Therefore, Sk is the required diatonic scale shift in S. 2.5.1 The solution is shown in the following gure:

Mathematics and Music, by James S. Walker.

69
S4 S1

S5

S2

S3

T S1 T FIGURE A.1. Five diatonic scale shifts in a Mozart passage, S4 , S5 , S3 , S5 , and S1 . There is also a succession of three note motives within the boxes marked S1 and S2, where the rst two notes of each motive are transposed (rst by S1 , then by S1 , and nally by S1 ). The notes in S2 are obtained from the diatonic scale shift S0 applied to S1, so we nearly have a hierarchy of diatonic scale shifts.
S5

2.5.3 The solutions to (a), (b), and (c) are shown in the following gure:

(a):

(b):

(c):

(d) The notes in (b) and (c) are different. 2.5.5 A solution is shown in Figure A.2.

P1

P3 P2

FIGURE A.2. Three palindromes, P1, P2, P3, in a Mozart passage.

Mathematics and Music, by James S. Walker.

70

APPENDIX A. SOLUTIONS TO ODD-NUMBERED EXERCISES

2.5.7 The solution is shown in the following gure:

2.5.9 R R reverses the notes twice, so it just repeats the original sequence of notes, and that is precisely what S0 does. 2.5.11 The original melody is the beginning of Frere Jacques (c.f. Figure 2.4). 2.6.1 2.6.3 E G We have G 7 E 4 C 0 so C-major chord 7 G-major chord . 2.6.5 2.6.7 2.6.9 2.6.11 2.6.13 The solution is shown in Figure A.3. E B A G 9 The chord is D-F-A. So we have C-major chord D-minor chord . The chord is A-C -E. So we have C-major chord 9 A-major chord . There are only 12 hours on the clock, so only 12 ways to shift pitches. We have T12+k = Tk , since 12 + k is the same shift as k on the 12-hour Chromatic Clock, so there can be no more than 12 distinct transpositions. Each Tk , for k = 0, 1, 2, . . . , 11, is distinct because Tk (0) = k.
T R T

B D G

A +7
+7

2 D 11 B 7 G

2.6.15 The solutions to each part are (a) Letting j be the hour number of the lowest pitch note, the hours for the notes in the major chord are j, j + 4, j + 7. A transposition Tk maps those hours to j + k, j + 4 + k, j + 7 + k, so the intervals remain 4 and 3 between low, middle, and high notes. (b) A minor chord has intervals of 3 and 4 between its low middle and high notes. Letting j be the hour number of the lowest pitch note, the hours for the notes in the major chord are j, j + 3, j + 7. A transposition Tk maps those hours to j + k, j + 3 + k, j + 7 + k, so the intervals remain 3 and 4 between low, middle, and high notes. (c) Letting j be the hour number of the lowest pitch note, the hours for the notes in the major chord are j, j + 4, j + 7. A chromatic inversion Rk maps those hours to k j, k j 4, k j 7. Writing these hours in reverse order, we have k j 7, k j 4, k j, which have intervals of 3 and 4. Those intervals specify a minor chord. On the other hand, for a major chord, we would have hours of j, j + 3, j + 7. The chromatic inversion Rk maps those hours to k j, k j 3, k j 7. Writing those hours in reverse order, we have k j 7, k j 3, k j, which have intervals of 4 and 3. Those intervals specify a major chord. (d) For any hour m, we have R0 Tk (m) = R0 Tk (m) = R0 m k = 0 (m k) = m + k = Tk (m) = Tk (0 m) = Tk R0 (m) = Tk R0 (m).
Mathematics and Music, by James S. Walker.

71

E
T0

T5

E
T 7

T5

T 5

E
C

FIGURE A.3: Opening measures of Ray Charles Whatd I Say, with transpositions indicated. The transformation indicated by C is a combination of the two separate transpositions, T7 and T5 .
Since R0 Tk (m) = Tk R0 (m) for each hour m, we have that R0 Tk = Tk R0 . We also have Rk (m) = k m, but we showed above that R0 Tk (m) = m + k. Therefore, Rk (m) = R0 Tk (m) for every hour m. Thus, Rk = R0 Tk . Combining the two results we just proved, yields Rk = Tk R0 . (e) Since a chromatic inversion Rk is equal to Tk R0 , the chromatic inversion R0 will map a given minor chord to a major chord, and then Tk will map that major chord to a major chord. Therefore, Rk maps every minor chord to a major chord. Similar reasoning shows tha Rk maps every major chord to a minor chord. (f) First, suppose both chords are major. The hours for the three notes of chord1 can be written as j, j + 4, j + 7, while those of chord2 can be written as , + 4, + 7. Then the transposition Tj maps chord1 to chord2 . Similar reasoning can be used if both chords are minor. Second, suppose that chord1 is minor and chord2 is major. If R0 is applied to chord1 then it becomes major, and we have shown that there is a transposition (call it Tk ), that maps this major chord to the major chord chord2 . Therefore, the composition Tk R0 maps chord1 to chord2 . We showed in part (d) that Tk = Tk R0 , so we have proved that the chromatic inversion Rk maps chord1 to chord2 . A similar argument applies the nal case of chord1 being major and chord2 being minor. 2.6.17 The semitone changes are 0, 3, 0, 1, 0, 1, 4, 11, 6, 11, 4, 1, 0, 1, 0, 3, 0 They exhibit a left/right symmetry about the middle number 6 with sign reversal. The rst half of the passage, ending at A , is retrograded and then transposed by +6 semitones to get the second half of the passage. In other
Mathematics and Music, by James S. Walker.

72

APPENDIX A. SOLUTIONS TO ODD-NUMBERED EXERCISES


words, the second half is obtained by applying the composition T6 R, where R is the retrograde transformation, to the rst half.

2.6.19 In the solution for Exercise 2.6.15(f), we showed that there is a transposition Tj that maps a given major chord chord1 to a given major chord chord2 . The argument given there clearly shows that Tj is unique. Likewise, the argument for the existence of a transposition mapping a given minor chord to a given minor chord given in the solution for Exercise 2.6.15(f), implies that that transposition is unique. If the given chord chord1 is minor, while the given chord chord2 is major, then the solution to Exercise 2.6.15(f) shows that the transformation that maps chord1 to chord2 has the form Rk . Suppose there was another transformation P that does the mapping. Then Rk P 1 maps the major chord chord2 to itself. But that transformation must then equal the unique transposition T0 . So we have the equation Rk P 1 = T0 and therefore Rk = T0 P. But T0 P = P, so we have Rk = P. That proves that Rk is unique. A similar argument proves uniqueness for a transformation mapping a given major chord to a given minor chord.

Chapter 3
3.1.1 The calculation goes as follows: r2 D
r2

r4 E

r2

r6 F

r1

r7 G

r2

r9 A

r2

r11 B

r2

r13 r1 C

r1

r2 D

3.1.3 A natural minor scale uses this pattern of additions on the Chromatic Clock: +2 +1 +2 +2 +1 +2 +2

and, because B has frequency multiplier r11 , we have r11 B


r2

r13 r1 C

r1

r2 D

r2

r4 E

r2

r6 F

r1

r7 G

r2

r9 A

r2

r11 B

3.1.5 The following table contains the answers to this exercise: Chord I: C4 -E4 -G4 ii: D4 -F4 -A4 iii: E4 -G4 -B4 IV: F4 -A4 -C5 V: G4 -B4 -D5 vi: A4 -C5 -E5 vii : B4 -D5 -F5 I: C5 -E5 -G5 Frequencies (Hz) 262, 330, 392 294, 349, 440 330, 392, 494 349, 440, 523 392, 494, 587 440, 523, 659 494, 587, 699 523, 659, 784 Multipliers r0 , r4 , r7 r2 , r5 , r9 r4 , r7 , r11 r5 , r9 , r12 r0 r7 , r11 , r14 r2 r9 , r12 r0 , r16 r4 r11 , r14 r2 , r17 r5 r12 r0 , r16 r4 , r19 r7

3.1.7 The frequency multiples for the original melody are r0 C r2 D r4 E r0 C r0 C r2 D r4 E r0 C r4 E r5 F r7 G

Mathematics and Music, by James S. Walker.

73
Since E has a frequency multiple of r4 , relative to C, we multiply all of the powers of r by r4 and write down the notes for these new powers of r: r4 E and the score looks like this: r6 F r8 G r4 E r4 E r6 F r8 G r4 E r8 G r9 A r11 B

3.1.9 Multiplying by r7 is transposing down a fth. Multiplying by r5 is transposing up a fourth, while multiplying by r5 is transposing down a fourth. Multiplying by r4 is transposing up a third, while multiplying by r4 is transposing down a third. Transposing up a sixth would be multiplying by r9 , while transposing down a sixth would be multiplying by r9 3.2.1 The following table shows the rst 6 harmonics for G2 and the notes they correspond to: 98 Hz G2 196 Hz G3 294 Hz D4 392 Hz G4 490 Hz B4 588 Hz D5

3.2.3 The following table shows the rst 6 harmonics for E2 and the notes they correspond to: 82 Hz E2 164 Hz E3 246 Hz B3 326 Hz E4 410 Hz G 4 486 Hz B4

3.2.5 The following table shows the 7 harmonics for C3 and the notes they correspond to: 131 Hz C3 262 Hz C4 393 Hz G4 524 Hz C5 655 Hz E5 786 Hz G5 917 Hz A 5

3.2.7 The half-period p/2 is shown in Figure A.4. The half-period p/2 has frequency 1/(p/2) = 2/p, which is 2 times 1/p = 329 Hz, so its frequency is 2 329 Hz. 16000 8000 0 8000 16000 0.012 0.023 0.035 0.046

p/2

FIGURE A.4. Waveform of portion of a ute note. The time p for an approximate cycle is p = 0.0304 seconds. The half-period p/2 is indicated. 3.3.1 The notes and consonances are shown in Table A.1. There are at least three points to note:
Mathematics and Music, by James S. Walker.

74

APPENDIX A. SOLUTIONS TO ODD-NUMBERED EXERCISES


1. There are 8 sets of consonances in Table A.1. These eight sets of consonances can be listed according to how they lie among the harmonics of the three notes in the chord, as follows (an asterisk denotes an harmonic that is not in consonance): (1): {C5 , , C5 } (5): {G6 , G6 , G6 } (2): {A5 , A5 , } (6): {A6 , A6 , } (3): {C6 , , C6 } (7): {C7 , , C7 } (4): {, E6 , E6 } (8): {E7 , E7 , E7 }.

2. The base note of the chord, C4 , has consonances with both of the other two notes of the chord, E4 and G4 . 3. There are several harmonics consonant with the notes B and D. 4. There are some harmonics that are not consonant with notes in the key of C-major, and some notes that are not consonant with any of the notes on the chromatic scale. Table A.1 P ITCHES OF H ARMONICS FOR F- MAJOR CHORD

F3 (F3 ): (F4 ): (C5 ): (F5 ): (A5 ): (C6 ): (D ): 6 (F6 ): (G6 ): (A6 ): (C7 ): 175 350 525 700 875 1050 1225 1400 1575 1750 1925 2100 2275 2450 (E7 ): (F7 ):

A3 (A3 ): (A4 ): (E5 ): (A5 ): (C ): 6 (E6 ): (G6 ): (A6 ): (B6 ): (C ): 7 (E7 ): 220 440 660 880 1100 1320 1540 1760 1980 2200 2420 2640

C4 (C4 ): (C5 ): (G5 ): (C6 ): (E6 ): (G6 ): (A ): 6 (C7 ): (D7 ): (E7 ): 262 524 786 1048 1310 1572 1834 2096 2358 2260

2625 2800

Based on Table 2.1. Notes in parentheses are on chromatic scale, consonances underlined.

3.3.3 Each major chord is a transposition of the C-major chord. For example, the chord F3 -A3 -C4 is created from the chord C4 -E4 -G4 by frequency multiplying by the factor r7 . The consonances in the chord C4 -E4 -G4 are (perceptually) matching frequency ratios for various harmonics. Consequently the corresponding harmonics in the chord F3 -A3 -C4 will also have (perceptually) matching frequency ratios. Similar remarks apply to the major chord G3 -B3 -D4 , which is a transposition from the chord C4 -E4 -G4 using the frequency multiplier r5 . 3.3.5 The notes and consonances are shown in Table A.2. There are at least three points to note: 1. There are 4 sets of consonances shown in Table A.2. These four sets of consonances can be listed according to how they lie among the harmonics of the three notes in the chord: (1): {B4 , , B4 } (2): {B5 , B5 , , B5 } (3): {F , , F } 6 6 (4): {B6 , B6 , B6 }.

Mathematics and Music, by James S. Walker.

75
2. In contrast to major chords, there are no consonances with harmonics of the middle note, G3 , of the chord. 3. There are 3 harmonics for G3 that are consonant with the note D. The chord E3 -G3 -B3 -D4 is a standard chord in music. It is an E-minor seventh chord. Table A.2 P ITCHES OF H ARMONICS FOR E- MINOR CHORD

E3 (E3 ): (E4 ): (B4 ): (E5 ): (G ): 5 (B5 ): 165 330 495 660 825 990 1155 (E6 ): (F ): 6 (G ): 6 (B6 ):

G3 (G3 ): (G4 ): (D4 ): (G5 ): (B5 ): (D6 ): 196 392 588 784 980 1176 1372 (G6 ): (A6 ): (B6 ): (D7 ): 1568 1764 1960 2156 2352

B3 (B3 ): (B4 ): (F ): 5 (B5 ): (D ): 6 (F ): 6 (B6 ): (C ): 7 247 494 741 988 1235 1482 1729 1976 2223

1320 1485 1650 1815 1980 2145

Based on Table 2.1. Notes in parentheses are on chromatic scale, consonances underlined.

3.3.7 Each minor chord is a transposition of the D-minor chord. For example, the chord E3 -G3 -B3 is created from the chord D4 -F4 -A4 by frequency multiplying by the factor r10 . The consonances in the chord D4 -F4 -A4 are (perceptually) matching frequency ratios for various harmonics. Consequently the corresponding harmonics in the chord E3 -G3 -B3 will also have (perceptually) matching frequency ratios. Similar remarks apply to the minor chord A3 -C4 -E4 , which is a transposition from the chord D4 -F4 -A4 using the frequency multiplier r5 . 3.3.9 The notes of the chord are D, A, F . If those chords are arranged as D, F , A, then the semitone changes on the chromatic clock are 4 and 3, so it is a major chord. The octaves for the notes are F , A4 , D5 , so it called a rst 4 inversion of a major chord (and it is being arpegiatted downward). 3.3.11 We calculate that r4 satises r4 = 24/12 = 1.25992104989487 which is closely approximate to 5/4. Since E4 has frequency C r4 , it follows that the frequency of E4 is closely approximated by (5/4) C . Consequently every fourth harmonic of E4 will be closely approximate to every fth harmonic of C4 . Since this argument only depends on r4 5/4, and does not depend on the value of C , it follows that the third note on a major scale (like E4 on the C-major scale) will have similar harmonies with the rst note on the major scale. 3.3.13 The fourth note on a major scale will be more harmonious than the third note. The reason is that every fourth harmonic of the starting note of the scale will be in consonance with a harmonic of the fourth note. But only every
Mathematics and Music, by James S. Walker.

76

APPENDIX A. SOLUTIONS TO ODD-NUMBERED EXERCISES


fth harmonic of the starting note of the scale will be in consonance with a harmonic of the third note. Therefore, the number of consonances (within the nite range of human hearing) is greater for the fourth note.

3.3.15 In Table 2.2 we see that the low note C4 has consonances with harmonics of each of the overtones of the higher note G4 . Omitting the middle note E4 from the standard triadic chord, eliminates its overtones G and G which could 6 7 be heard as dissonant with overtones of G4 . 3.3.17 (a) If the two successive notes have frequencies of 1 and 2 , then 2 /1 = r = 21/12 . Therefore the (cents) measure of their difference is (cents) = 1200 log2 (2 /1 ) = 1200 log2 (21/12 ) 1 = 1200 12 = 100. (b) We have (cents) = 1200 log2 (3951/3930) = 1200 log(3951/3930)/ log(2) = 9.22624059804212 . . . and that is where we got 9.23 cents. 3.3.19 The notes and consonances are shown in Table A.3. The consonances are exactly the same as those discussed in Example 2.3.2 for the equal-tempered scale (compare with Table 2.3). The difference between the two cases, is that the notes in the just scale are slightly lower in pitch and the relative closeness of consonant notes will differ from those found in the equal-tempered case (sometimes closer for just scale, while sometimes closer for equal-tempered scale). 3.3.21 The transposed C-major scale and the correct B-major scale are shown in the following table, along with the cents differences between the frequencies for the notes on the two scales. Note B C D E F G A B B-major (Just Scale) 480 546 614 640 717 819 896 960 Transposed C-major 480 540 600 639.38 720 800.63 900 960 (Cents) Difference 0 19.13 39.93 1.69 7.23 39.28 7.71 0

The notes that differ the most are D and G . For those notes, the transposed notes are about 40 cents below the values on the B-major scale. So they are nearly a half semitone at.

Mathematics and Music, by James S. Walker.

77

Table A.3

P ITCHES OF H ARMONICS FOR D- MINOR CHORD

D4 (D4 ): (D5 ): (A5 ): (D6 ): (F ): 6 (A6 ): (D7 ): (E7 ): (F ): 7 (A7 ):

F4 288 576 864 (F4 ): (F5 ): (C6 ): (F6 ): (A6 ): (C7 ): (F7 ): (G7 ): (A7 ): (C8 ): 341 682 1023 1364 1705 2046 2387 2728 3069 3410 3751 4092

A4 (A4 ): (A5 ): (E6 ): (A6 ): (C ): 7 (E7 ): (A7 ): (B7 ): 427 854 1281 1708 2135 2562 2989 3416 3843

1152 1440 1728 2016 2304 2592 2880 3168 3456 3744

Based on Table 2.8. Notes in parentheses are on chromatic scale, consonances underlined.

Mathematics and Music, by James S. Walker.

You might also like