Jump to content

Pulse wave: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
tweaks
slight tweaks
Line 6: Line 6:
[[File:PWM duty cycle with label.gif|thumb|Duty cycles]]
[[File:PWM duty cycle with label.gif|thumb|Duty cycles]]


A '''pulse wave''' or '''pulse train''' or '''rectangular wave''' is a [[non-sinusoidal]] [[waveform]] that is the [[Periodic function|periodic]] version of the [[rectangular function]]. It is held high a percent each cycle ([[Period of a function|period]]) called the [[duty cycle]] and for the remainder of the cycle is low. A duty cycle of 50% produces a [[square wave]]s while other duty cycles result in asymmetrical rectangular waves. The average level of a rectangular wave is also given by the duty cycle.
A '''pulse wave''' or '''pulse train''' or '''rectangular wave''' is a [[non-sinusoidal]] [[waveform]] that is the [[Periodic function|periodic]] version of the [[rectangular function]]. It is held high a percent each cycle ([[Period of a function|period]]) called the [[duty cycle]] and for the remainder of each cycle is low. A duty cycle of 50% produces a [[square wave]], a specific case of a rectangular wave. The average level of a rectangular wave is also given by the duty cycle.


[[Pulse-width modulation]] varies the duty cycle.
[[Pulse-width modulation]] refers to methods that vary the duty cycle of a pulse wave.


==Frequency-domain representation==
==Frequency-domain representation==

Revision as of 20:43, 19 February 2024

The shape of the pulse wave is defined by its duty cycle d, which is the ratio between the pulse duration (τ) and the period (T).
Duty cycles

A pulse wave or pulse train or rectangular wave is a non-sinusoidal waveform that is the periodic version of the rectangular function. It is held high a percent each cycle (period) called the duty cycle and for the remainder of each cycle is low. A duty cycle of 50% produces a square wave, a specific case of a rectangular wave. The average level of a rectangular wave is also given by the duty cycle.

Pulse-width modulation refers to methods that vary the duty cycle of a pulse wave.

Frequency-domain representation

The Fourier series expansion for a rectangular pulse wave with period , amplitude and pulse length is[1] where .

Equivalently, if duty cycle is used, and :

Note that, for symmetry, the starting time () in this expansion is halfway through the first pulse.

Alternatively, can be written using the Sinc function, using the definition , as or with as

Generation

A pulse wave can be created by subtracting a sawtooth wave from a phase-shifted version of itself. If the sawtooth waves are bandlimited, the resulting pulse wave is bandlimited, too.

Fourier series of a 33.3% pulse wave, first fifty harmonics (summation in red)

Applications

The harmonic spectrum of a pulse wave is determined by the duty cycle.[2][3][4][5][6][7][8][9] Acoustically, the rectangular wave has been described variously as having a narrow[10]/thin,[11][3][4][12][13] nasal[11][3][4][10]/buzzy[13]/biting,[12] clear,[2] resonant,[2] rich,[3][13] round[3][13] and bright[13] sound. Pulse waves are used in many Steve Winwood songs, such as "While You See a Chance".[10]

In digital electronics, a digital signal is a pulse train (a pulse amplitude modulated signal), a sequence of fixed-width square wave electrical pulses or light pulses, each occupying one of two discrete levels of amplitude.[14][15] These electronic pulse trains are typically generated by metal–oxide–semiconductor field-effect transistor (MOSFET) devices due to their rapid on–off electronic switching behavior, in contrast to BJT transistors which slowly generate signals more closely resembling sine waves.[16]

See also

References

  1. ^ Smith, Steven W. The Scientist & Engineer's Guide to Digital Signal Processing ISBN 978-0966017632
  2. ^ a b c Holmes, Thom (2015). Electronic and Experimental Music, p.230. Routledge. ISBN 9781317410232.
  3. ^ a b c d e Souvignier, Todd (2003). Loops and Grooves, p.12. Hal Leonard. ISBN 9780634048135.
  4. ^ a b c Cann, Simon (2011). How to Make a Noise, [unpaginated]. BookBaby. ISBN 9780955495540.
  5. ^ Pejrolo, Andrea and Metcalfe, Scott B. (2017). Creating Sounds from Scratch, p.56. Oxford University Press. ISBN 9780199921881.
  6. ^ Snoman, Rick (2013). Dance Music Manual, p.11. Taylor & Francis. ISBN 9781136115745.
  7. ^ Skiadas, Christos H. and Skiadas, Charilaos; eds. (2017). Handbook of Applications of Chaos Theory, [unpaginated]. CRC Press. ISBN 9781315356549.
  8. ^ "Electronic Music Interactive: 14. Square and Rectangle Waves", UOregon.edu.
  9. ^ Hartmann, William M. (2004). Signals, Sound, and Sensation, p.109. Springer Science & Business Media. ISBN 9781563962837.
  10. ^ a b c Kovarsky, Jerry (Jan 15, 2015). "Synth Soloing in the Style of Steve Winwood". KeyboardMag.com. Retrieved May 4, 2018.
  11. ^ a b Reid, Gordon (February 2000). "Synth Secrets: Modulation", SoundOnSound.com. Retrieved May 4, 2018.
  12. ^ a b Aikin, Jim (2004). Power Tools for Synthesizer Programming, p.55-56. Hal Leonard. ISBN 9781617745089.
  13. ^ a b c d e Hurtig, Brent (1988). Synthesizer Basics, p.23. Hal Leonard. ISBN 9780881887143.
  14. ^ B. SOMANATHAN NAIR (2002). Digital electronics and logic design. PHI Learning Pvt. Ltd. p. 289. ISBN 9788120319561. Digital signals are fixed-width pulses, which occupy only one of two levels of amplitude.
  15. ^ Joseph Migga Kizza (2005). Computer Network Security. Springer Science & Business Media. ISBN 9780387204734.
  16. ^ "Applying MOSFETs to Today's Power-Switching Designs". Electronic Design. 23 May 2016. Retrieved 10 August 2019.