musical sound

musical sound

Introduction

      any tone with characteristics such as controlled pitch and timbre. The sounds are produced by instruments in which the periodic vibrations can be controlled by the performer.

      That some sounds are intrinsically musical, while others are not, is an oversimplification. From the tinkle of a bell to the slam of a door, any sound is a potential ingredient for the kinds of sound organization called music. The choices of sounds for music making have been severely limited in all places and periods by a diversity of physical, aesthetic, and cultural considerations. This article will analyze those involved in Western musical traditions.

      The fundamental distinction usually made has been between tone and noise, a distinction best clarified by referring to the physical characteristics of sound. Tone differs from noise mainly in that it possesses features that enable it to be regarded as autonomous. Noises are most readily identified, not by their character but by their sources; e.g., the noise of the dripping faucet, the grating chalk, or the squeaking gate. Although tones too are commonly linked with their sources (violin tone, flute tone, etc.), they more readily achieve autonomy because they possess controlled pitch, loudness, timbre, and duration, attributes that make them amenable to musical organization. Instruments that yield musical sounds, or tones, are those that produce periodic vibrations (vibration). Their periodicity is their controllable (i.e., musical) basis.

      The strings of the violin, the lips of the trumpet player, the reed of a saxophone, and the wooden slabs of a xylophone are all, in their unique ways, producers of periodic vibrations. The pitch, or high-low aspect, created by each of these vibrating bodies is most directly a product of vibrational frequency. timbre (tone colour) is a product of the total complement of simultaneous motions enacted by any medium during its vibration. loudness is a product of the intensity of that motion. Duration is the length of time that a tone persists.

 Each of these attributes is revealed in the wave (wave motion) form of a tone. The pattern may be visualized as an elastic reed—like that of a clarinet—fixed at one end, moving like a pendulum in a to-and-fro pattern when set into motion (see illustration—>). Clearly, this reed's motion will be in proportion to the applied force. Its arc of movement will be lesser or greater depending upon the degree of pressure used to set it into motion. Once moving, it will oscillate until friction and its own inertia cause it to return to its original state of rest. As it moves through its arc the reed passes through a periodic number of cycles per time unit, although its speed is not constant. With these conditions prevailing, its motion through time could be charted by placing a carbon stylus on its moving head, then pulling a strip of paper beneath it at a uniform rate. The reed's displacement to-and-fro diminishes in a smooth fashion as time passes (decreasing intensity). Each cycle of its arc is equally spaced (uniform frequency). Each period of the motion forms the same arc pattern (uniform wave content). If this vibratory motion were audible, it could be described as follows: it grows weaker from the beginning (diminishing loudness) until it becomes inaudible; it remains at a stable level of highness (steady pitch); and it is of unvarying tonal quality (uniform timbre). If the reed were a part of a clarinet and the player continued blowing it with unvaried pressure, loudness, pitch, and timbre would appear as constants.

Tone
      Most musical tones differ from the demonstration tone (above) in that they consist of more than a single wave form. Any material undergoing vibratory motion imposes its own characteristic oscillations on the fundamental vibration. The reed probably would vibrate in parts as well as a whole, thus creating partial wave forms in addition to the fundamental wave form. These partials are not fortuitous. They bear harmonic relationships to the fundamental motion that are expressible as frequency ratios of 1:2, 3:4, etc. This means that the reed (or string or air column as well) is vibrating in halves and thirds and fourths as well as a whole. Another way of expressing this is that half the body is vibrating at a frequency twice as great as the whole; a third is vibrating at a frequency three times greater; etc.

 These numerical relationships also are expressible by pitch relationships as the harmonic, or overtone, series (see illustration—>), which is merely a representation of numerical ratios in terms of pitch equivalents. Depending upon its shape and substance, a vibrating mass performs motions that are the equivalents of these partial vibrations, whether it be the mass of a string, reed, woodblock, or air column. This means that most tones are composites: they consist of partial vibrations of the vibrating body as well as the vibrations of the whole mass. Although one can develop the acuity required to hear some of these overtones within a musical tone, the ear normally ignores them as separate parts, recognizing only a more or less rich tone quality within the fundamental pitch.

      Although pure tones, or tones lacking other than a fundamental frequency, sometimes occur in music, most musical tones are composites. A typical violin tone is relatively rich in overtones while a flute tone sometimes approaches a pure tone. What the listener recognizes as “a violin tone” or “a trumpet tone” also is a product of the noise content that accompanies the articulation of any sound on the particular instrument. The friction of the bow as it is set into motion across the string, the eddies of air pressure within a horn's mouthpiece, or the hammer's impact on a piano string all add an extra dimension, a significant “noise factor,” to any manually produced tone. After articulation, however, it is the presence or absence of overtones and their relative intensities that determine the timbre of any tone. The violin and flute tones are distinguishable because their articulatory “noises” are quite different and their overtone contents are dissimilar, even when they produce the same pitch.

      Musical tones of determined harmonic content can be produced by electronic vacuum tubes (electron tube) or transistors as well as by traditional manual instruments. Some electronic organs (electronic organ), for example, use single vacuum tubes whose frequency output can be varied through control of an adjustable transformer. Through ingenious mixing circuits a compound tone consisting of any predetermined overtone content can be produced, thereby imitating the sound of any traditional instrument. Composers of electronic music have utilized this capability to synthesize tones quite different from any available on traditional instruments, as well as tones similar to natural sounds. Electronic computers (computer music) are capable of complete imitation of such sounds; the tone is broken down into its component parts, then synthesized through an auditory output circuit.

Movement
      Once an audible oscillation is produced by a vibrating body, it moves away from its source as a spherical pressure wave. Its rate of passage through any medium is determined by the medium's density and elasticity; the denser the medium, the slower the transmission; the greater the elasticity, the faster. In air at around 60° F, sound moves at approximately 1,120 feet per second, the rate increasing by 1.1 feet per second per degree of rise in temperature.

      Sound waves move as a succession of compressions through the air. The wavelength is determined by frequency; the higher the pitch, the shorter the wavelength. A pitch of 263 cycles per second (middle C of the piano) is borne as a wavelength of around 4.3 feet (speed of sound ÷ frequency = wavelength).

      By the time a wave has moved some distance, it has changed in some of its characteristics. The journey has robbed it of intensity, which is inversely proportional to the square of the distance. Its timbre has been altered slightly by objects within its path that disrupted an equitable distribution of frequencies, particularly the high-frequency waves, which, unlike the low, move in relatively straight paths from their sources.

      The area within which a sound occurs can have considerable effect upon what is heard. Just as a string or reed or air column has a natural resonance period (or rate of vibration (resonator)), any enclosure—whether an audio speaker cabinet or the nave of a cathedral—imposes its resonance characteristics on a sound wave within it. Any tone that approximates in frequency the characteristic resonance period of an enclosure will be reinforced through the sympathetic response, or natural resonance, of the air within the enclosure. This means that tones of frequencies differing from the resonance of the enclosure will be less intense than those that agree, thereby creating an inequity of sound intensities.

      Fortunately, most rooms where music is performed are large enough (wall lengths greater than about 30 feet) so that their natural resonance periods are too slow to fall within the range of pitches of the lowest musical tones (usually no lower than 27 cycles per second, although some organs have pipes that extend to 15 cycles per second). Smaller rooms can produce disturbing sympathetic resonance unless obstructions or absorbent materials are added to minimize that effect. (Bathroom singers revel in this phenomenon because the band of resonance sometimes lies close enough to the pitches of the male voice to support it, making it appear richer and more powerful.)

      In addition to resonance, any enclosure possesses a reverberation period, a unit of time measured from the instant a sound fills the enclosure (steady state) until that sound has decayed to one-millionth of its initial intensity. Anyone who has spoken or clapped his hands inside a large, empty room has experienced prolonged reverberation. There are two reasons for such protracted reverberation: first, the space between the surfaces of the enclosure is so great that reflected sound waves travel extended distances before decaying; and, second, the absence of highly absorbent materials precludes appreciable loss of intensity of the wave during its movement.

      The reverberation period is a crucial factor in rooms where sounds must be heard with considerable fidelity. If the period is too long in a room where speech must be understood, spoken syllables will blend into each other and the words will be mumbled confusion. If, on the other hand, the reverberation period is too brief in a room where human “presence” and music each contribute to the acoustics (acoustics, architectural), only a “cold” and “dull” feeling will persist, because no reverberative support of the prevailing sounds can be provided by the enclosure itself. (See also acoustics: Architectural acoustics (acoustics).)

      Although all sound waves, regardless of their pitch, travel at the same rate of speed through a particular medium, low tones mushroom out in a broad trajectory while high tones move in straight paths. For this reason listeners in any room should be within a direct path of sound propagation. Seats far to the side at the front of an auditorium offer occupants a potentially distorted version of sound from its source. Thus the high-frequency speakers (loudspeaker) (tweeters) in good audio reproduction systems are angled toward the sides of the room, ensuring wider coverage for high-frequency components of all sounds.

      Sites of musical performance in the open demand quite different acoustical arrangements, of course, since sound reflection from ceilings and walls cannot occur and reverberation cannot provide the desirable support that would be available within a room. A reflective (mirror) shell placed behind the sound source can provide a boost in transmission of sounds toward listeners. Such a reflector must be designed so that relatively uniform wave propagation will reach all locations where listening will occur. The shell form serves that purpose admirably since its curved shape avoids the right angles that might set up continuous reflections, or echoing. Furthermore, sound waves are reflected more uniformly over a wide area than with any other shape, diffusing them equally over the path of propagations. (The needs here are similar to those of the photographer who wishes to flood a scene uniformly with flat light rather than focus with a spotlight on a small area.)

Pitch and timbre
      Just as various denominations of coins combine to form the larger units of a monetary system, so musical tones combine (musical composition) to form larger units of musical experience. Although pitch, loudness, duration, and timbre act as four-fold coordinates in the structuring of these units, pitch has been favoured as the dominating attribute by most Western theorists. The history of music theory has to a great degree consisted of a commentary on the ways pitches are combined to make musical patterns, leaving loudness and timbre more as the “understood” parameters of the musical palette.

      Music terminology, for example, recognizes loudnesses in music in terms of an eight-level continuum of nuances from “extremely soft” (ppp, or pianississimo) to “extremely loud” (fff or fortississimo). (The musical dominance of Italy from the late 16th to the 18th century—when these Italian terms first were applied—explains their retention today.)

      The timbres of music enjoy an even less explicit and formalized ranking; other than the vague classifications “shrill,” “mellow,” “full,” and so on, there is no standard taxonomy of tone quality. Musicians for the most part are content to denote a particular timbre by the name of the instrument that produced it.

Division of the pitch spectrum
      Pitch is another matter. A highly developed musical culture demands a precise standardization of pitch, and Western theory has been occupied with this task from as early as Aristoxenus (4th century BC). Especially since the Renaissance, when instruments emerged as the principal vehicles of the musical impulse, problems of pitch location (tuning) and representation (notation) have challenged the practicing musician. When at least two instrumentalists sit down to play a duet, there must be some agreement about pitch, or only frustration will result. Although the standardization of the pitch name a′ (within the middle of the piano keyboard) at 440 cycles per second has been adopted by most of the professional music world, there was a day—even during the mid-18th century of Bach—when pitch uniformity was unknown.

      Man's perception (hearing) of pitch is confined within a span of roughly 15 to 18,000 cycles per second. This upper limit varies with the age and ear structure of the individual, the upper limit normally attenuating with advancing age. The pitch spectrum is divided into octaves (octave), a name derived from the scale theories of earlier times when only eight (Latin octo) notes within this breadth were codified. Today the octave is considered in Western music to define the boundaries for the pitches of the chromatic scale. The piano keyboard is a useful visual representation of this 12-unit division of the octave. Beginning on any key, there are 12 different keys (and thus 12 different pitches), counting the beginning key, before a key occupying the same position in the pattern recurs.

      One must keep in mind that the chromatic scale, within the various octave registers of man's hearing, is merely a conventional standard of pitch tuning. Performers like singers, trombone and string players, who can alter the pitches they produce, frequently make use of pitches that do not correspond precisely to this set of norms. The music of many non-Western cultures also utilizes distinct divisions of the octave. Furthermore, some contemporary music makes use of pitch placements that divide the octave into units smaller than the half-step. This music, called microtonal (microtonal music), has not become standard fare in Western cultures, in spite of its advocates (Alois Hába, Julian Carillo, Karlheinz Stockhausen) and even its special instruments that provide a means for consistent performance.

      Western music history is dotted with systems formulated for the precise tuning (tuning and temperament) of pitches within the octave. From a modern viewpoint all suffer from one of two mutually exclusive faults: either they lack relationships (intervals (interval)) of uniform size, or they are incapable of providing chords (chord) that are acceptable to the ear. Pythagorean tuning provides uniformity but not the chords. Just tuning (just intonation), based on the simpler ratios of the overtone series, provides the chords but suffers from inequality of intervals. Meantone (meantone temperament) tuning provides equal intervals but gives rise to several objectionable chords, even in simple music. All three of these systems fail to provide the pitch wherewithal for the 12 musical keys found in the standard repertoire.

      The compromise tuning system most widely accepted since the mid-19th century is called “ equal temperament.” Based on the division of the octave into 12 equal half-steps, or semitones, this method provides precisely equal intervals and a full set of chords that, although not as euphonious as those of the overtone series, are not offensive to the listener.

 The semitone is the smallest acknowledged interval of the Western pitch system. The sizes of all remaining intervals can be calculated by determining how many semitones each contains. The names of these intervals are derived from musical notation through a simple counting of lines and spaces of the staff (see illustration—>). Just as the overtone content of a single tone determines timbre, the relationship of the constituent pitches of an interval determines its quality, or sonance. There is a long history of speculations in this area, but the subjectivity of the data indicates that little verifiable fact can be sorted from it.

Consonance and dissonance
      Until the 20th century, music theorists were prone to concoct tables that showed an “objective” classification of intervals into the two opposing camps of consonant (consonance and dissonance) and dissonant. But only the person who utters these terms can know with assurance what he means by them, although many attempts have been made to link consonant with pleasant, smooth, stable, beautiful and dissonant with unpleasant, grating, unstable, and ugly. These adjectives may be reasonably meaningful in musical contexts, but difficulty arises if one attempts to pin a singular evaluation on a particular interval per se.

      Theorists have noted that the character of an interval is altered considerably by the sounds that surround it. Thus the naked interval that sounds “grating,” “unstable,” and lacking in fusion might within a particular context create an altogether different effect, and vice versa.

      Recognition of the power of context in shaping a response to the individual pitch interval has led some music theorists to think more in terms of a continuum of sonance that extends from more consonant to more dissonant, tearing down the artificial fence once presumed to separate the two in experience.

 The explanation of consonance and dissonance offered by Hermann von Helmholtz (Helmholtz, Hermann von) in On the Sensations of Tone (1863) is perhaps as helpful as any. An initial theory was based on the notion that dissonance is a product of beats, which result from simultaneous tones or their upper overtones of slightly differing frequencies. Another explanation, offered later by Helmholtz, held that two tones are consonant if they have one or more overtones (excluding the seventh and ninth) in common (see illustration—>).

      Music in which a high degree of dissonance occurs has rekindled interest in this old problem of psychoacoustics. The German composer Paul Hindemith (Hindemith, Paul) (1895–1963) provided one explanation of harmonic (harmony) tension and relaxation that depends upon the intervals found within chords. According to his view a chord is more dissonant than another if it contains a greater number of intervals that, as separate entities, are dissonant. Although Hindemith's reasonings and conclusions have not been widely accepted, the absence of any more convincing explanation and classification often leads musicians to use his ideas implicitly.

      Although the complete pitch spectrum can be tuned in a way that provides 12 pitches per octave (as the chromatic scale), pitch organization in music usually is discussed in terms of less inclusive kinds of scale patterns. The most important scales in traditional Western theory are seventone (heptatonic (heptatonic scale)), which, like the chromatic, operate within the octave. These scales are different from one another only in the intervals formed by their constituent pitches. The major scale, for instance, consists of seven pitches arranged in the intervallic order: tone–tone–semitone–tone–tone–tone–semitone.

      Called major because of the large (or major) third that separates the first and third pitches, this scale differs from the minor scale mainly in that the latter contains a small (or minor) third in this location. Since three variants of the minor scale are recognized in the music of the Western repertoire, it is important to note that they share this small interval between their first and third pitches.

Scales and modes
      Major and minor scales formed the primary pitch ingredients of music written between 1650 and 1900, although this is a sweeping generalization for which exceptions are not rare. Other scales, called modes (mode), possess greater representational power for music of earlier times and for much of the repertoire of Western folk music. These too are heptatonic patterns, their uniqueness produced solely by the differing pitch relationships formed by their members. Each of the modes can most easily be reproduced by playing successive white keys at the piano.

      The modes and the major and minor scales best represent the pitch structure of Western music, though they do not utilize the total complement of 12 chromatic pitches per octave. They are abstractions that are meaningful for tonal (tonality) music; i.e., music in which a particular pitch acts as a focal point of perception, establishing a sense of repose or tonality to which the remaining six pitches relate. Major and minor scale tonality was basic to Western music until it began to disintegrate in the art music of the late 19th century. It was replaced in part by the methods of Arnold Schoenberg (Schoenberg, Arnold) (1874–1951), which used all 12 notes (serialism) as basic material. Since that revolution of the early 1920s, the raw pitch materials of Western music have frequently been drawn from the complete chromatic potential. By contrast, the music of several Eastern cultures, a number of children's songs, and occasional Western folk songs incorporate pitch materials best classified as pentatonic (pentatonic scale) (a five-pitch scale).

Sound production of musical instruments (musical instrument)
      Excluding electronic tone synthesizers, which employ vacuum tubes or transistors to produce tones, musical instruments can be classified within three groups: (1) chordophones, or strings; (2) aerophones, or winds; and (3) idiophones and membranophones, nearly all of which are percussion instruments. Each category is further divisible into groups according to the way the vibrating medium is set into motion.

Chordophones (chordophone)
      Three means of eliciting sounds determine three categories within the family of chordophones. They are bowing, plucking, and striking. Most common of the first category are the violin, viola, violoncello, and double bass of the orchestra, all of which use a horsehair bow for setting their strings into motion. Essentially a resonant box bearing strings (stringed instrument) of four different fundamental frequencies, members of this group have not changed appreciably in construction since the 17th century, except for the 20th-century advent of the electrified bass, which is in fact a close cousin of the amplified guitar.

      Violins and the larger members of its group are sounded by plucking (pizzicato) on occasion, which provides a brittle tone of extremely brief duration. The harp is the best known orchestral instrument whose tone depends upon the noise components added by plucking. Other plucked instruments are the guitar, banjo, mandolin, ukelele, zither, lyre, lute, and the harpsichord. The latter differs from the piano in that its strings are actuated by the plucking action of a tiny plectrum.

      The piano is most notable of the struck stringed instruments, employing a hammer mechanism linked with the keyboard for producing its wide range of sounds. Other instruments of this group are the clavichord and the dulcimer.

      For all chordophones pitch is proportional to string tension and inversely proportional to length, thickness, and density. Since string length is the most readily altered of these factors, all chordophones provide a means for altering the resonating length of strings (as with the violin and guitar) or a set of many string lengths and masses (as with the piano and harp) for producing a variety of pitches.

Aerophones (aerophone)
      This category covers everything from the piccolo to the pipe organ and is best understood by consistent reference to the nature of the air (wind instrument) column employed in the various types of instruments, as well as the way this air column is set into motion.

      Brass instruments (brass instrument) consist of a long tube whose cross section is proportionately small. Coupled with a mouthpiece that, in response to vibrations of the performer's lips, helps to create eddies of air pressure that set an enclosed air column into motion, these instruments produce a range of pitches corresponding to the overtone series. The bugle is a primitive kind of brass instrument in that it is limited to only one overtone series, while the modern trumpet, cornet, French horn, trombone, tuba, flügelhorn, and various kinds of euphoniums utilize valves (valve) or a slide to lengthen the air column and thus provide up to seven different overtone series. Pitch on these instruments is primarily a function of tube length, the wavelength of the instrument's fundamental pitch equal to twice the length of the tube, plus a so-called end correction that accommodates variations of bore. Timbre is a product of mouthpiece shape, bore (whether cylindrical or conical), and material, aside from the important role performed by the player himself in obtaining desired overtones.

      Woodwinds (woodwind) prior to the 20th century were made for the most part of wood. Today the flutes and clarinets are classified in this group only because of this heritage, while the saxophones, always built of metal, share only the reed (reed instrument) mouthpiece and similar fingering technique with the clarinet. All are, nonetheless, called woodwinds, and they consist of an air column set into motion by one of two means: (1) through high pressure eddies produced by the wind of the performer blown directly into the instrument (as with a recorder or whistle) or over it (as with the flute and piccolo), or (2) by means of a vibrating reed that is set into motion by air pressure from the performer. The clarinets and saxophones utilize a single reed fixed at one end, while the oboe, English horn, and bassoon use two thin reeds that are connected laterally and vibrate jointly. For all of these instruments, either keys or the fingers of the performer directly open holes, with the effect of shortening the enclosed air column of the instrument and thereby producing higher fundamental pitches. Through overblowing and various fingering procedures, the overtone series provides the wealth of pitches available on these instruments.

      Free reed instruments utilize a single, freely vibrating reed, different in nature from that of a woodwind. The category includes the accordion, harmonica, and harmonium and their relatives. In these instruments the reed vibrates, causing periodic vibrations in the air; but the reed's size, rather than the air enclosed by the instrument, determines the pitch.

      Pipe organs (organ) are of the aerophone (wind) category, too, although their keyboard mechanism and literature link them closely with the piano and harpsichord. Like a grand synthesis of woodwinds and brasses, organs produce their tones by means of tuned air columns that are formed with pipes of varied length, cross section, and shape (called flue pipes) or by means of a vibrating brass reed actuated by forced air (called reed pipes). Flue pipes range in length from under an inch to 32 feet.

Idiophones (idiophone) and membranophones (membranophone)
      Idiophones are instruments whose bodies vibrate to produce sound. The class contains most of the pitched percussion instruments (percussion instrument). These include instruments made of wood or other organic material, such as xylophones. They also include pitched percussioninstruments that are struck or plucked and are made of metal or other inorganic material (triangle, glockenspiel, vibraphone, celesta, tubular bell, gong, steel drum, cymbal, glass harmonica, etc.). Idiophones without pitch consist of such instruments as the percussion board, castanets, and rattles, all of which are made of wood or other organic material and are struck, scraped, rubbed, brushed, or shaken.

      Membranophones produce sound by a vibrating membrane. The group consists most notably of the timpani, or kettledrums, which can be tuned by increasing or decreasing the tension of the membranes that form the heads of the enclosed cavities. Other membranophones consist of drums without fixed pitch, such as side drums, bongos, and various non-Western types of fixed and indefinite pitch. Tone quality and character are the result of the player's skill in controlling intensity and overtone character of the sound.

William E. Thomson Ed.

Additional Reading
Willem A. Van Bergeijk, J.R. Pierce, and E.A. David, Waves and the Ear (1960); Arthur H. Benade, Horns, Strings and Harmony (1960); Charles A. Culver, Musical Acoustics, 4th ed. (1956); John Backus, The Acoustical Foundations of Music (1969); John Mills, A Fugue in Cycles and Bels (1935); Max F. Meyer, How We Hear: How Tones Make Music (1950); Llewelyn S. Lloyd, The Musical Ear (1940); Alexander Wood, The Physics of Music, 6th ed. (1962); René Descartes, Musicae compendium (1650; Eng. trans., Compendium of Music, 1961); Cornelis J. Nederveen, Acoustical Aspects of Woodwind Instruments (1969).

* * *


Universalium. 2010.

Игры ⚽ Нужен реферат?

Look at other dictionaries:

  • Sound and music computing — (SMC) is a research field that studies the whole sound and music communication chain from a multidisciplinary point of view. By combining scientific, technological and artistic methodologies it aims at understanding, modeling and generating sound …   Wikipedia

  • sound — sound1 soundable, adj. /sownd/, n. 1. the sensation produced by stimulation of the organs of hearing by vibrations transmitted through the air or other medium. 2. mechanical vibrations transmitted through an elastic medium, traveling in air at a… …   Universalium

  • musical instrument — Introduction       any device for producing a musical sound. The principal types of such instruments, classified by the method of producing sound, are percussion (percussion instrument), stringed (stringed instrument), keyboard (keyboard… …   Universalium

  • musical performance — Introduction       step in the musical process during which musical ideas are realized and transmitted to a listener. In Western music, performance is most commonly viewed as an interpretive art, though it is not always merely that. A performer… …   Universalium

  • musical notation — Written, printed, or other visual representation of music. There are two basic approaches to notating music. Tablature (such as guitar chord diagrams) depicts the actions a performer is to take (in particular, showing where to put the fingers to… …   Universalium

  • Musical gesture — Tonic and dominant in C  Play (help …   Wikipedia

  • Musical gestures — Movement associated with music, either physical (e.g. body movement) or mental (e.g. musical imagery) are musical gestures. The concept of musical gestures has received much attention in various disciplines studying music (e.g. musicology, music… …   Wikipedia

  • sound — 1. n. & v. n. 1 a sensation caused in the ear by the vibration of the surrounding air or other medium. 2 a vibrations causing this sensation. b similar vibrations whether audible or not. 3 what is or may be heard. 4 an idea or impression conveyed …   Useful english dictionary

  • sound — sound1 noun 1》 vibrations which travel through the air or another medium and are sensed by the ear.     ↘a thing that can be heard.     ↘the area within which something can be heard. 2》 (also musical sound) sound produced by continuous and… …   English new terms dictionary

  • musical*/ — [ˈmjuːzɪk(ə)l] adj I 1) involving music, or relating to music musical instruments[/ex] an evening of musical entertainment[/ex] 2) a musical sound is pleasant to listen to 3) good at playing or singing music They re a very musical family.[/ex]… …   Dictionary for writing and speaking English

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”