You are on page 1of 11

Dictionary of the History of Ideas

MUSIC AND SCIENCE

OF ALL the arts, music has had most to rely upon a scientific and mathematical analysis of its materials. Both the relations between pitches and between durations are best defined by numbers and ratios. To construct even the simplest instruments out of strings or pipes, musicians had to derive as best they could the laws of sound production. The most elementary fact, generally accredited to the Pythagoreans but probably known to the ancient Babylonians and Egyptians, is that if a string is stopped in the middle, each of the two halves sounds an octave higher than the whole; if divided into three parts, two-thirds of the string will sound a fifth above the whole; and so on. Because it relies on precise measurement, music was considered until fairly modern times, indeed until around 1650, a branch of science. In late antiquity it began to be included in the four mathematical disciplines of the quadrivium along with arithmetic, geometry, and astronomy. But actually only theoretical music was accorded this place. No singing or playing was included in this curriculum. Practical music making went its own way, maintaining only limited contact with theoretical music, drifting farthest from it in the Middle Ages and approaching nearer during the Renaissance. The musical component of the mathematical curriculum in the universities never went beyond the heritage of Greek music theory. Only the Renaissance humanists succeeded in making this relevant to Western musical art. Because of this alliance with mathematics, music figured prominently in cosmology, astrology, and number mysticism. Speculations about the harmony of the universe were often inspired by musical facts, as in Plato's Timaeus (31-39), or as in the theory that the planets were governed in their motions by ratios of the consonances and therefore produced an unheard music (Republic X). These ideas were attacked by Aristotle (On the Heavens II. 8-9), but the musical world generally believed them until the end of the

fifteenth century, and Kepler much later was still seeking to prove universal harmony when he discovered the third law of planetary motion (the cube of any planet's distance from the sun varies directly with the square of the planet's period or time of revolution). Greek writers credited Pythagoras (ca. 582-07 B.C.) with the earliest acoustical observations. He is said to have discovered the ratios of the octave (2:1), fifth

261

(3:2), octave plus fifth (3:1), fourth (4:3), and double octave (4:1). These were the only consonances recognized by Greek theory. Great metaphysical significance was attached to the fact that the set of numbers from 1 to 4 was the source of all harmony. It was assumed that these ratios produced the same consonances whether the numbers applied to string lengths, bore of pipes, weights stretching strings, weights of disks, or volumes of air in vessels such as bells or water-filled glasses. Theon of Smyrna (second century A.D.) claimed that Pythagoras had verified these ratios in all these circumstances. Boethius (fifth century A.D.) reported that Pythagoras heard the consonances also issuing from a blacksmith's hammers whose weights were in the same ratios as the string lengths. Actually, as Vincenzo Galilei, father of Galileo, was to demonstrate in 1589 (Palisca, 1961), the ratios are not the same in these cases as in the divison of the

string. Throughout the Middle Ages and early Renaissance, from Boethius to Gaffurio (Figure 1), almost every author on music recounts the experiments of Pythagoras without realizing their improbability. The canonization of the octave, fifth, and fourth in their natural ratios as the cornerstones of the harmonic system had a deeper influence on music theory and composition than on instrument building or playing. Musicians tended to tune their instruments by ear, tempering the ratios of the fourths and fifths, because it was discovered early that if one tuned up a cycle of twelve fifths from any note until that note was reached again, this note was higher than that reached through a cycle of seven octaves by ( 3/2 )12 : (2/1)7 or 531,441/524,288 approximately 24% of a semitone. According to the theorists, the only acceptable tuning was that which maintained the fifth and fourth at their proper ratios of 3:2 and 4:3. This tuning was called by Ptolemy Diatonic ditoniaion, and it is also known as the Pythagorean tuning. In this scheme the tetrachord or modular fourth is composed of two tones and a semitone in the ratios 9:8, 9:8, 256:243. The fifth and fourth, favored by this tuning, were the most prominent consonances in written polyphony from the ninth through the thirteenth centuries, particularly at points of rhythmic or structural emphasis. The thirds and sixths, which were not recognized as consonances by Greek or orthodox medieval theory, were harshsounding in the Pythagorean tuning. Nevertheless, they were employed increasingly in polyphony, particularly during the fourteenth and fifteenth centuries. Renaissance humanism had a somewhat delayed effect on musical theory as compared to other disciplines. Only in the last quarter of the fifteenth century did the ancient Greek treatises begin to be read firsthand and translated; for example, Ptolemy's Harmonics, Aristotle's Problems, the short introductions of Cleonides and Euclid, and eventually the Harmonics of Aristoxenus. At about the same time an antitheoretical movement began among composer-theorists. The first writer to break with the Boethian-Pythagorean doctrine of consonances was Bartolom Ramos de Pareja. In his Musica practica (Bologna, 1482) he proposed a tuning system that allowed for sweetly tuned

thirds in the simple ratios of 5:4 and 6:5, as against the Pythagorean 81:64 and 32:27. Ramos' innovations met resistance among conservatives like Franchino Gaffurio. But soon Lodovico Fogliano (1529), Gioseffo Zarlino (1588), and Francisco de Salinas (1577) joined Ramos in dethroning for all times the Pythagorean tuning system. All three leaned upon the recently rediscovered Harmonics of Claudius Ptolemy in which a tuning very similar to Ramos' was described under the name Diatonic syntonon. Zarlino was convinced

262

that it was the perfect tuning, because both perfect and imperfect consonances were in simple ratios of the class n + 1/n, known as superparticular. Zarlino modernized the pre-Ramos number mysticism by replacing the number four by the set of numbers from 1 to 6, the numero senario. A growing use of thirds and sixths paralleled theoretical recognition of a sweet-third tuning. However, if there is a causal connection, it is that the theorists saw the anachronism of standing by a theoretical harsh-third tuning. The astronomer Ptolemy was known to the Renaissance as the theorist who took the middle road between the rationalist position of the Pythagoreans and the empiricist method of Aristoxenus. Zarlino was attracted to Ptolemy because he too was inclined to worship

number while aiming to satisfy the ear. Aristoxenus, on the other hand, rejected ratios as irrelevant to music. He preferred to divide pitch-space directly, if somewhat subjectively. Of his tuning systems the one that most appealed to sixteenth-century musicians was that in which each whole tone contained 12 units of pitchspace and each semitone 6 units. Sixteenth-century interpreters assumed this meant the division of the octave into an equal temperament of twelve equal semitones, as in the tuning of the lute. Such a tuning would permit a melody to sound equally well in any key, something that could not be accomplished with any other tuning. Aristoxenus began to find apostles in the last quarter of the sixteenth century, notably Vincenzo Galilei and Ercole Bottrigari. Meanwhile no scientific discovery had yet deprived the simple ratios of the consonances of their priority. But in 1589-90 Vincenzo Galilei drafted a treatise that reported some new experiments with sounding bodies. He discovered that the ratios usually associated with the consonances are obtained only when they represent pipe or string lengths, other factors being equal. When weights were attached to strings, the ratio had to be 4:1, not 2:1 to produce an octave. The volumes of concave bodies had to be in the ratio of 8:1 to produce the octave. Since, in terms of weights, the fifth and fourth were 9:4 and 16:9 respectively, Galilei saw no reason to prefer simple ratios within the numbers of the senario. The bastion of the simple ratios was besieged also by another line of research. In a letter to the composer Cipriano de Rore of around 1563 the scientist Giambattista Benedetti proposed a new theory of the cause of consonance. Benedetti argued that since sound consists of air waves or vibrations, in the more consonant intervals the shorter more frequent waves concurred with the longer less frequent waves at regular intervals. In the less consonant intervals, on the other hand, concurrence was infrequent and the two sounds did not blend in the ear pleasantly. He showed that in a fifth, for example, the two vibrations will meet every two cycles of the lower note and every three of the higher. He went on to show that in terms of frequency of concurrence the hierarchy of ratios within the octave would be 2:1, 3:2, 4:3, 5:3, 5:4, 6:5, 7:5,

8:5, which challenges both the superiority of superparticular ratios and the sanctity of the senario. There could be no abrupt break from consonance to dissonance but only a continuum of intervals, some more, some less consonant. Benedetti's theory was espoused in the next century by Isaac Beeckman and Marin Mersenne, who sought Ren Descartes' opinion of it. Descartes declined to judge the goodness of consonances by such a rational method, protesting that the ear prefers one or another according to the musical context rather than because of any concordance of vibrations. The philologist and student of ancient music Girolamo Mei summed up this emancipation of music from scientific determinism in a letter to Vincenzo Galilei of 1572:
The true end of science is altogether different from that of art.... The science of music goes about diligently investigating and considering all the qualities and properties of the existing constitution and ordering of musical tones, whether these are simple qualities or comparative, like the consonances, and this for no other aim than to come to know the truth itself, the perfect goal of all speculation, and as a by-product the false. It then lets art exploit as it sees fit, without any limitation, those tones about which science has learned the truth

(Palisca [1960], p. 65). The revolution in musical thought encouraged experimentation in composition, in which a search for new musical resources had already spontaneously begun. Composers found a new harmonic richness; and even in the old tunings they braved modulations to distant keys and ventured melodic motion by semitone. The Aristoxenian equal temperament, which would have made these things easier, was demanded even by some conservative musicians like Giovanni Maria Artusi, a loyal disciple of Zarlino. Dissonances seconds, sevenths, and diminished and augmented intervalswere introduced more and more freely into compositions. If scientific discovery stimulated musical change, the opposite is also true: musical problems stimulated

scientific investigation. Benedetti and Vincenzo Galilei were moved by musical problems to inquire into the mechanics of sound production. The most notable case is that of Galileo, who was disturbed by the very problem that stumped his father: Is there a stable

263

connection between consonance and ratio? He made perhaps the most fundamental discovery in acoustics when he proved that there is: ratios between the frequencies of vibration are the inverse of the ratios of string lengths. The most difficult challenge that music presented to science in the seventeenth century was to explain the multiple pitches that could be heard when a single string vibrated. Aristotle noted that he could hear the octave above (Problems 919b 24; 921b 42) and observed the related phenomenon of hearing a string respond sympathetically to one tuned an octave higher. It took Mersenne's acute musical ear to hear from a single vibrating string not only the upper octave but the octave plus fifth, double octave, double octave plus major third, and the double octave plus major sixth. Neither her nor Descartes, nor any of the other scientists of their circle could explain why this happened. In 1673 two Oxford scientists, William Noble and Thomas Pigot, showed that strings tuned to the octave, octave plus fifth, and octave plus major third below a plucked string sounded sympathetically at the unison to the

plucked string by vibrating in aliquot parts. They demonstrated this by placing paper riders on the sympathetic strings at the points where, if the string were stopped, unisons would be produced. In 1677 John Wallis reported that multiple sounds would occur in a vibrating string only if it was not plucked at the points that marked off the aliquot parts. This showed that a single string simultaneously vibrated as a whole and in its aliquot parts. Thus harmonic vibration, which was important also for mechanics and optics, was established as a fact. Of all the laws of acoustics that of harmonic vibration exercised most the imagination of theorists of music. The first to utilize the information was JeanPhilippe Rameau. In his Trait de l'harmonie (1722), he had constructed a new system of harmony on the ratios of the divisions of the string. When he learned rather belatedly of harmonic vibration from an exhaustive paper by Joseph Sauveur (1701), Rameau decided that this was the original principle he had been looking for. In his opinion it firmly established his theory of fundamental bass, as he called the bottom note of a triad whose notes are arranged in thirds, for the first six notes of the harmonic series arranged as a chord is equivalent to a major triad over a fundamental bass. Thus his system was a copy of nature. The fundamental bass, in his view, determines the progress of the harmony, as when it leaps down a fifth from the dominant (fifth note) to the tonic (first note) of a key. But Rameau did not stop at this. He made of the harmonic series a Cartesian first principle from which he built up, by manipulating the numbers of its ratios, a system of theory that embraced every aspect of music. Unfortunately, his numerical operations were often faulty and drew severe criticism from the geometer Jean d'Alembert and the mathematician Leonhard Euler. The concept of fundamental bass received further support when the celebrated violinist and composer Giuseppe Tartini in 1754 announced his discovery of the third sound. This is a subjective sensation now known as difference tone that is believed to occur because of the presence of nonlinear resonance in the ear. When two pitches are sounded, a third lower one seems to resound. Tartini found it by listening carefully to double-stops played on a violin. Actually, unknown

to Tartini, Georg Andreas Sorge had noted the same phenomenon nine years previously (1745). The third sound usually reinforced the note of a chord that Rameau identified as the fundamental bass, which Tartini too accepted as a keystone of his system. Like Rameau, he indulged in sweeping mathematical and geometric speculations, which, however, did not withstand the scrutiny of mathematicians such as Benjamin Stillingfleet and Antonio Eximeno. Both Tartini and Rameau were pioneers in musical composition, at the forefront in technique, style, and structure. But as thinkers they were anachronistic. In the midst of the Enlightenment, which was skeptical of systems and the deductive process, they spun webs of numbers in blissful isolation, only to become hopelessly entangled in them. Musicians have continued to search for a natural basis for music theory. In the twentieth century, Paul Hindemith extended the idea of fundamental bass to all kinds of dissonant chords. Like his predecessors he based the theory on the harmonic series and on difference tones. He believed, like Rameau, that all harmonic movement depends on the progress of roots of chords, but he freed this process from the simple cadence-like successions of Rameau. Hindemith's theory has been challenged because its premisses are fully valid only in a system of just intonation, and because, while exploiting difference tones, he ignored combination tones that are sometimes more audible. Lately, scientific facts and theories outside the realm of acoustics have inspired philosophies of music. From thermodynamics the concepts of entropy and indeterminacy have been seized upon as a justification for music to copy nature by following laws of chance rather than willful combinations. From information theory and physics composers have borrowed the concept of the stochastic process, in which events are interconnected through a succession of probabilities. It is also possible by analogy to defend as expressive of the contemporary view of reality the modular structures of serial compositions, which are constructed like

264

crystals, multiplying the same cell in successive and simultaneous juxtapositions. Meanwhile the entire corpus of acoustical science is called into play by electronic and computer music, in which art merges indissolubly with engineering. Music may be on the road to becoming once again a branch of science, or at least of technology.
BIBLIOGRAPHY

For references concerning Beeckman, Benedetti, Descartes, Fogliano, V. Galilei, G. Galilei, Gaffurio, Kepler, Mersenne, Mei, B. R. de Pareja, Rameau, Salinas, and Sauveur, see C. V. Palisca, Scientific Empiricism in Musical Thought, in H. H. Rhys, ed., Seventeenth-Century Science and the Arts (Princeton, 1961), pp. 91-137. See also Jean Le Rond d'Alembert, lmens de musique, thorique et pratique (Lyon, 1762); J. M. Barbour, Tuning and Temperament (East Lansing, Mich., 1953); E. Bottrigari, Il Desiderio (Venice, 1594); M. R. Cohen and I. E. Drabkin, A Source Book in Greek Science (New York, 1948), pp. 286-310; R. L. Crocker, Pythagorean Mathematics and Music, The Journal of Aesthetics and Art Criticism, 22 (1963-64), 189-98, 325-35; Signalia Dostrovsky, The Origins of Vibration Theory: The Scientific Revolution and the Nature of Music (Ph.D. dissertation, Princeton University, 1969, unpubl.); Stillman Drake, Renaissance Music and Experimental Science, Journal of the History of Ideas, 30 (1969), 483-500; L. Euler, Tentamen novae theoriae musicae (Petropoli [Saint Petersburg], 1739); A. Eximeno, Dell'origine delle regole (Rome, 1774); E. E. Helm, The Vibrating String of the Pythagoreans, Scientific American, 217 (1967),

92-103; H. Helmholtz, On the Sensations of Tone (New York, 1964); P. Hindemith, Craft of Musical Composition (New York, 1942); E. Lippman, Musical Thought in Ancient Greece (New York, 1964); C. V. Palisca, Girolamo Mei (Rome, 1960); idem, The Interaction of the Sciences and the Arts: A Historical View, Proceedings of the Fourth National Conference on the Arts in Education (Philadelphia, 1965), pp. 19-25; idem, Fogliano, Galilei, Gogava, Mei, Ramos, Salinas, Valla, Zarlino, Die Musik in Geschichte und Gegenwart (Kassel, 1955-68); G. A. Sorge, Vorgemach der musikalischen Composition, Erster Theil (Lobenstein, 1745), Ch. 5; B. Stillingfleet, Principles and Power of Harmony (London, 1771); G. Tartini, Trattato di musica (Padua, 1754). CLAUDE V. PALISCA [See also Astrology; Cosmology; Number; Pythagorean Harmony; Renaissance Humanism.]

You might also like