Saturday 8 March 2014

OFC 008. KISWAHILI ----- FOUNDATION COURSE -- THE OPEN UNIVERSITY OF TANZANIA---- BY. MWL JAPHET MASATU.

OFC  OO8 . KISWAHILI ----  FOUNDATION  COURSE ----BY.  MWL . JAPHET  MASATU.

Maana ya Fasihi


.1:  MAANA YA FASIHI

Kuna waandishi tofautitofauti ambao wamejitahidi kuelezea juu ya maana au fasili ya fasihi kama ifuatavyo;
Fasihi; ni sanaa yaani mkusanyiko wa kazi mbalimbali zilizosanifiwa kwa kutumia lugha itumiayo zaidi maneno na kujishughulisha na jinsi binadamu anavyojitambua mwenyewe binafsi,Pia anavyoathiriwa na binadamu wenzake aidha na viumbe vingine aina kwa aina katika mazingira mbalimbali ya maisha.(Nkwera, Fr.F.V,2003:94)
Fasihi; ni sanaa ya lugha inayoshughulikia masuala yanayomhusu binadamu matatizo yake ,ndoto zake,matumaini yake,migogoro yake na jinsi anavyoingiliana na mazingira yake.(Wamitila,2004:19).
Fasihi; ni sanaa itoayo maudhui yake kwa kutumia lugha ya maneno ambayo hutamkwa ama kuandikwa.(Senkoro F.E.M.K,2011:11).
Kwa mtazamo wangu  kutokana na maana za waanishi tofautitofauti ambao wamejitahidi kutoa maana ya fasihi mimi nimeona kuwa fasihi ni kazi ya sanaa inayotumia lugha katika njia  masimulizi au maandishi katika kufikisha ujumbe uliokusudiwa  kwa jamii husika katika kuikabili mazingira yake katika nyanja mbalimbali za maisha kamavile siasa uchumi, na utamaduni
Ndani ya fasihi kama sanaa kuna matawi au tanzu mbili za fasihi nazo ni fasihi simulizi na fasihi andishi. Katika


1.2:   M AANA YA FASIHI ANDISHI

Fasihi andishi; hii ni fasihi iliyozuka baada ya kuvumbuliwa kwa mfumo wa maandishi. Huhifadhiwa kwenye maandishi, ili yasomwe na vizazi vya baadaye. (Wamitila K.W,2004) .
Pia wahusika katika fasihi Andishi fasihi hutumia wahusika kama wanadamu, kinyume na fasihi Simulizi ambayo hutumia wahusika changamano (watu, wanyama, mazimwi n.k). Kunazo aina kadhaa za wahusika katika kazi ya fasihi andishi kama(http://www.gafkosoft.com/sw/search/?s=fasihi).


1.3:   MAANA YA FASIHI SIMULIZI

Fasihi simulizi ni kazi ya sanaa  inayotumia lugha, Kazi hii ya sanaa huhifadhiwa kwa  kichwa na kusambazwa kwa njia ya masimulizi ya mdomo. (Nchimbi.A.S,2011).
Fasihi simulizi ni kazi ya sanaa inayotumia mdomo (tofauti na maandishi kama kwenye fasihi andishi) kama njia kuu maalumu kwa ajili ya kufikisha ujumbe. Ni sehemu ya msingi ya utamaduni, na vipengere vyake vingi vinafanya kazi kama katika fasihi kwa ujumla. (htt://sw.m.wikipedia.org/wiki/fasihi simulizi)

Fasihi ya Kiafrika


Makala hii ina dalili ya kutungwa kwa kutegemea programu ya kompyuta kama "google translation kit". Watumiaji wanaombwa kuchungulia lugha, viungo na muundo wake tena. Wakiridhika na hali yake wanaweza kuondoa kigezo hiki kinachoonekana kwenye dirisha la kuhariri juu ya matini ya makala kwa alama za {{tafsiri kompyuta}} .
Fasihi ya Kiafrika inamaanisha maandiko kuhusu na kutoka Afrika. Kama George Joseph anavyodokeza katika ukurasa wa kwanza wa sura yake kuhusu Fasihi ya Afrika katika kitabu understanding Contemporary Africa, ' wakati mtizamo wa fasihi ya Ulaya kwa ujumla inahusu herufu zilizoandikwa, dhana ya Afrika inahusisha pia fasihi simulizi. [1]
Kama George Joseph anavyoendelea kuandika, wakati maoni ya Ulaya kuhusu fasihi yalisisitiza mgawanyo wa sanaa na maudhui, mwamko wa Afrika unajumuisha:
"Fasihi" inaweza pia kuashiria matumizi ya maneno kisanaa kwa ajili ya sanaa pekee. Bila kukana jukumu muhimu la somo la sanaa katika Afrika, tunapaswa kukumbuka kwamba, tangu jadi, Waafrika huwa hawatenganishi sanaa na kufundisha. Badala ya kuandika au kuimba kwa sababu ya uzuri ulioko, Waandishi wa Afrika, kwa kufuata fasihi simulizi, hutumia uzuri kusaidia kuwasiliana ukweli na taarifa muhimu na jamii. Hakika, kitu huchukuliwa kizuri kwa sababu ya ukweli kinachoonyesha na jamii kinachosaidia kujenga. [2]

Maandiko ya kitambo

(Misri ya Kale) ulikuwa miongoni mwa tamaduni za kale zaidi za Kiafrika zenye desturi ya kuandika, ambao baadhi ya maandishi yake yaliyotumia picha na alama kuashiria kitu au jambo huishi mpaka waleo. Kazi kama vile Kitabu cha Wafu cha Misri kawaida hutumiwa na wasomi kama rejeleo la msingi la imani ya kidini ya Misri ya kale na sherehe. Maandiko ya Kinubi, yanakubalika lakini kwa wakati huu hayasomeki. Awali yaliandikwa kwa kutumia picha na alama na hatimaye kwa kutumia alfabeti yenye miundo 23, kutafsiri kumekuwa kugumu.


Fasihi simulizi

Fasihi simulizi inaweza kuwa katika nathari au ushairi. Mara nyingi, nathari huwa hadithi au historia na inaweza kujumuisha hekaya zenye wahusika walaghai. Wasimuliza katika Afrika wakati mwingine hutumia mbinu ya kuuliza-na-majibu kusimulia hadithi zao. Mashairi, mara nyingi huimbwa, hujumuisha: masimulizi ya visa, aya ya shughuli, aya ya kidini, mashairi ya kusifu watawala na watu wengine maarufu. Waimbaji wa sifa[[]], husimulia hadithi zao kwa kuimba. [3] Nyimbo za upendo, za kazi, za watoto, misemo, mithali na vitendawili pia husomwa au huimbwa. [4]

Fasihi kabla ya Ukoloni

Mifano ya fasihi ya Kiafrika kabla ya ukoloni ni mingi. Fasihi simulizi ya Afrika ya magharibi inajumuisha visa vya Sundiata vilivyotungwa katika Mali ya kitambo, Visa vya zamani vya Dinga kutoka Dola la kitambo la Ghana. Nchini Ethiopia, iliyoandikwa awali kwa miundo ya Ge'ez ni Kebra Negast au kitabu cha wafalme. Aina moja maarufu ya ngano za kiafrika ni hadithi za "ulaghai", ambapo wanyama wadogo hutumia akili zao kushinda mapambano na wanyama wakubwa. Mifano wa wanyama laghai ni pamoja na Anansi, buibui katika ngano za Waashanti kutoka Ghana; Ijàpá, kobe katika ngano za Wayoruba kutoka Nigeria; na Sungura, ambaye hupatikana katika ngano za Afrika ya Mashariki na kati. [5] Kazi zingine zilizoandikwa ni tele, yaani katika Afrika kaskazini, kanda ya Sahel ya Afrika magharibi na pwani ya Kiswahili. Kutoka Timbuktu peke yake, kuna wastani nakala 300,000 au zaidi zilizowekwa katika maktaba mbalimbali na mikusanyiko ya kibinafsi,[6] hasa zilizoandikwa kwa Kiarabu, lakini baadhi yazo katika lugha za wenyeji (yaani Peul na Songhai) [7] Nyingi ziliandikiwa katika Chuo Kikuu maarufu cha Timbuktu. Kazi hizi zinaongelea mada nyingi, ambazo ni pamoja na Unajimu, Ushairi, Sheria, Historia, Imani, Siasa na Falsafa miongoni mwa zingine. [8] Fasihi ya Kiswahili vile vile, imehamasishwa na mafundisho ya Kiislam lakini ilikua katika mazingira ya kienyeji. Mojawapo wa vipande vya fasihi ya Kiswahili vinavyojulikana na vya kitambo sana ni Utendi wa Tambuka au "Hadithi ya Tambuka". Katika nyakati za Kiislamu, Waafrika wa Magharibi kama vile ibn Khaldun walitia fora sana katika fasihi ya Kiarabu. Afrika Kaskazini ya kitambo ilinufaisha Vyuo vikuu kama vile vya Fez na Cairo, kwa kiasi kikubwa mno cha maandiko ili kuongeza yao.


Fasihi ya Kiafrika wakati wa ukoloni

Kazi nzuri za Kiafrika zinazojulikana Ulaya kutoka kipindi cha ukoloni na kile cha biashara ya utumwa ni za masimulizi ya utumwa, kama vile The Interesting Narrative of the Life of Olaudah Equiano (1789) cha Olaudah Equiano.

Katika kipindi cha ukoloni, Waafrika walioathiriwa na lugha za Ulaya walianza kuandika katika lugha hizo. Mwaka 1911, Yusufu Efraimu Casely-Hayford (ajulikanaye pia kama Ekra-Agiman) kutoka Pwani ya Dhahabu (sasa Ghana) alichapisha ile pengine ni riwaya ya kwanza ya Kiafrika kundikwa kwa Kimombo, Ethiopia Unbound: Studies in Race Emancipation [9] Ingawa kazi hii iligusia hadithi na utetezi wa kisiasa, uchapishaji na kupitiwa upya na wanahabari wa Ulaya kulimaanisha wakati muhimu katika fasihi ya Kiafrika.


Katika kipindi hiki, tamthlia za Kiafrika zilianza kuibuka. Herbert Isaka Ernest Dhlomo wa Afrika ya Kusini alichapisha tamthlia ya kwanza ya Kiafrika kwa Kiingereza, [10] katika mwaka wa 1935. Mwakani 1962, Ngugi wa Thiong'o kutoka Kenya aliandika mchezo wa kuigiza wa kwanza wa Afrika ya Mashariki, The Black Hermit, hadithi iliyohadhari "ukabila" (ubaguzi kati ya makabila ya Afrika).


Fasihi ya Afrika mwishoni mwa kipindi cha ukoloni (kati ya mwisho wa Vita Vikuu vya Kwanza Vya Dunia na uhuru) iliendelea kuonesha mandhari ya ukombozi, uhuru, na (kati ya Waafrika kwenye maeneo yaliyodhibitiwa na Ufaransa ) kutambua weusi wa Waafrika. Mmoja wa viongozi wa harakati za kutambua weusi, mshairi na hatimaye Rais wa Senegal, Sédar Léopold Senghor, alichapisha mkusanyiko wa kwanza wa mashairi ya lugha ya Kifaransa yaliyoandikwa na Waafrika mwaka 1948, Anthologie de la Nouvelle poésie nègre et malgache de langue Française (Anthology of the New Black and Malagasy Poetry in French Language), akishirikisha utangulizi kutoka kwa mwandishi wa Kifaransa, Jean-Paul Sartre. [10]


Sio kwamba waandishi wa fasihi ya Afrika wakati huu walikuwa mbali na masuala waliyoyakabili. Wengi, kwa hakika, waliteswa kwa undani na pia kwa kuelekezwa: Akilaaniwa kwa kuyaweka kando majukumu yake ya kisanaa ili kushiriki kikamilifu katika mapambano, Christopher Okigbo aliuawa katika vita ya Biafra dhidi ya wanaharakati wa Nigeria katika vita vya wenyewe kwa wenyewe vya miaka ya 1960; Mongane Wally Serote aliwekwa kizuizini chini ya sheria ya Afrika Kusini ya Terrorism Act No 83 ya 1967 kati ya 1969 na 1970, na hatimaye kutolewa bila hata kujibu mashtaka; mjini London mwaka 1970, mwananchi mwenzake Arthur Norje alijiua; Jack Mapanje wa Malawi aliwekwa kizuizini bila kesi au mashtaka kwa sababu ya matamshi aliyotoa katika baa ya chuo kikuu ; na, mwakani 1995, Ken saro-Wiwa alikufa kwa kutiwa kitanzi na serikali ya wanajeshi ya Nigeria.

Fasihi ya Afrika baada ya ukoloni

Kwa ukombozi na kuongezeka kwa kusoma na kuandika kwani mataifa mengi ya Afrika yalipata uhuru wao katika miaka ya 1950 na 1960, fasihi ya Afrika imeongezeka kwa kasi kwa wingi na katika kutambuliwa, kwa kazi nyingi za Afrika kuonekana katika mitaala ya kielimu ya Ulaya na katika orodha ya "bora ya" vilivyoandikwa mwishoni mwa karne ya 20. Waandishi wa Afrika katika kipindi hiki waliandika katika lugha za Ulaya (hasa Kiingereza, Kifaransa na Kireno) na pia lugha za Afrika za jadi.

Ali A. Mazrui na wengine hutaja migogoro saba kama mandhari: Mfarakano kati ya Afrika iliyopita na iliyopo, kati ya Utamaduni na Usasa, kati ya wazawa na kigeni, kati ya kujijali na kujali jamii, kati Ujamaa na Ubepari, kati ya maendeleo na kujitegemea na kati Uafrika na ubinadamu. [11] Mandhari mengine katika kipindi hiki ni pamoja na matatizo ya kijamii kama vile ufisadi, tofauti za kiuchumi kati ya nchi ambazo zimepata uhuru maajuzi, na haki na majukumu ya wanawake. Waandishi wa kike sasa wamewakilishwa vyema katika fasihi ya Afrika iliyochapishwa kuliko walivyokuwa kabla ya uhuru.


Mwaka 1986, Wole Soyinka alikuwa mwandishi wa Afrika wa kwanza baada ya uhuru kushinda Tuzo la Nobel katika fasihi. Camus Albertmzaliwa wa Algeria alipokea tuzo hilo katika mwaka 1957.

Tuzo la Noma

Tuzo la Noma, lilianza mwaka wa 1980, hupatianwa kwa kazi bora ya mwaka katika fasihi ya Afrika.

Riwaya Kuu za Afrika

Washairi wakuu wa Afrika



Fasihi ya sekondari

  • kamusi elezo ya fasihi ya Kiafrika, ed Simon Gikandi, London: Routledge, 2003.
  • Historia ya Cambridge kuhusu fasihi ya Afrika na Caribbean, ed Abiola Irele na Simoni Gikandi, 2 vls, Cambridge [ua]: Cambridge University Press, 2004. Yaliyomo
  • Mabinti wa Afrika: Mkusanyiko wa Kimataifa wa Maneno na Maandishi wa ukoo wa Wanawake wa Afrika ", ed Margaret Busby (Random House, 1992).
  • Historia jumla ya Afrika. VIII, ed. Ali A. Mazrui, UNESCO, 1993, ch. 19 "fasihi ya kisasa tangu 1935," Ali A. Mazrui et al.
  • Kuelewa Afrika ya Kisasa, ed. Aprili A. Gordon na Donald L. Gordon, Lynne Rienner, London, 1996, ch. 12 "Fasihi ya Afrika", George Joseph

Angalia Pia

Lango la Fasihi

Marejeo

  1. Jump up [0] ^ George, Yusufu, "Fasihi ya Afrika" ch. 12 of Kuelewa Afrika ya kisasa s. 303
  2. Jump up ibid s. 304
  3. Jump up http://www.infoplease.com/ce6/ent/A0802673.html
  4. Jump up George Joseph, op. cit. uk. 306-310
  5. Jump up African Literature - MSN Encarta. Jalada kutoka ya awali juu ya 2009-10-31.
  6. Jump up http://www.sum.uio.no/research/mali/timbuktu/project/timanus.pdf
  7. Jump up http://www.spiegel.de/international/world/0, 1518,569560,00. Html
  8. Jump up http://www.loc.gov/exhibits/mali/
  9. Jump up [9] ^ Fasihi ya Afrika.
  10. Jump up Leopold Senghor - MSN Encarta. Jalada kutoka ya awali juu ya 2009-10-31.
  11. Jump up Ali A. Mazrui et al. "Maendeleo ya kisasa ya fasihi tangu 1935" kama ch. 19 ya UNESCO ya Historia Kuu ya Afrika ujazo. VIII s. 564f Ushirikiano na Ali A. Mazrui juu ya sura hii walikuwa Mario Pinto de Andrade, M'hamed Alaoui Abdalaoui, Daniel P. Kunene na Jan Vansina.

Viungo vya nje


OPH 111 . FUNDAMENTAL PHYSICS ---BY. MWL. JAPHET MASATU

Fundamentals of Physics


Fundamentals of Physics


The cover page of Fundamentals of Physics Extended 9th edition.
Authors David Halliday, Robert Resnick, Jearl Walker
Country United States of America
Language American English
Subject Physics
Genre Textbook
Published 1960 (John Wiley & Sons, Inc.)
Media type Print (hardcover)
Fundamentals of Physics is a calculus-based physics textbook by David Halliday, Robert Resnick, and Jearl Walker. [1] The textbook is currently in its tenth edition and is published in a five-volume set. The current version is a revised version of the original textbook Physics by Halliday and Resnick, first published in 1960. It is widely used in colleges as part of the undergraduate physics courses, and has been well known to science and engineering students for decades as "the gold standard" of freshman-level physics texts. In 2002, the American Physical Society named the work the most outstanding introductory physics text of the 20th century.
The textbook covers most of the basic topics in physics:
The extended edition also contains introductions to topics such as quantum mechanics, atomic theory, solid-state physics, nuclear physics and cosmology. A solutions manual and a study guide are also available. [
From Wikipedia, the free encyclopedia

Mechanics (Greek Μηχανική) is the branch of science concerned with the behavior of physical bodies when subjected to forces or displacements, and the subsequent effects of the bodies on their environment. The scientific discipline has its origins in Ancient Greece with the writings of Aristotle and Archimedes[1][2][3] (see History of classical mechanics and Timeline of classical mechanics). During the early modern period, scientists such as Galileo, Kepler, and especially Newton, laid the foundation for what is now known as classical mechanics. It is a branch of classical physics that deals with particles that are either at rest or are moving with velocities significantly less than the speed of light. It can also be defined as a branch of science which deals with the motion of and forces on objects.


Classical versus quantum

The major division of the mechanics discipline separates classical mechanics from quantum mechanics.
Historically, classical mechanics came first, while quantum mechanics is a comparatively recent invention. Classical mechanics originated with Isaac Newton's laws of motion in Principia Mathematica; Quantum Mechanics was discovered in 1925. Both are commonly held to constitute the most certain knowledge that exists about physical nature. Classical mechanics has especially often been viewed as a model for other so-called exact sciences. Essential in this respect is the relentless use of mathematics in theories, as well as the decisive role played by experiment in generating and testing them.
Quantum mechanics is of a wider scope, as it encompasses classical mechanics as a sub-discipline which applies under certain restricted circumstances. According to the correspondence principle, there is no contradiction or conflict between the two subjects, each simply pertains to specific situations. The correspondence principle states that the behavior of systems described by quantum theories reproduces classical physics in the limit of large quantum numbers. Quantum mechanics has superseded classical mechanics at the foundational level and is indispensable for the explanation and prediction of processes at molecular and (sub)atomic level. However, for macroscopic processes classical mechanics is able to solve problems which are unmanageably difficult in quantum mechanics and hence remains useful and well used. Modern descriptions of such behavior begin with a careful definition of such quantities as displacement (distance moved), time, velocity, acceleration, mass, and force. Until about 400 years ago, however, motion was explained from a very different point of view. For example, following the ideas of Greek philosopher and scientist Aristotle, scientists reasoned that a cannonball falls down because its natural position is in the Earth; the sun, the moon, and the stars travel in circles around the earth because it is the nature of heavenly objects to travel in perfect circles.
The Italian physicist and astronomer Galileo brought together the ideas of other great thinkers of his time and began to analyze motion in terms of distance traveled from some starting position and the time that it took. He showed that the speed of falling objects increases steadily during the time of their fall. This acceleration is the same for heavy objects as for light ones, provided air friction (air resistance) is discounted. The English mathematician and physicist Isaac Newton improved this analysis by defining force and mass and relating these to acceleration. For objects traveling at speeds close to the speed of light, Newton’s laws were superseded by Albert Einstein’s theory of relativity. For atomic and subatomic particles, Newton’s laws were superseded by quantum theory. For everyday phenomena, however, Newton’s three laws of motion remain the cornerstone of dynamics, which is the study of what causes motion.

Relativistic versus Newtonian mechanics

In analogy to the distinction between quantum and classical mechanics, Einstein's general and special theories of relativity have expanded the scope of Newton and Galileo's formulation of mechanics. The differences between relativistic and Newtonian mechanics become significant and even dominant as the velocity of a massive body approaches the speed of light. For instance, in Newtonian mechanics, Newton's laws of motion specify that F=ma, whereas in Relativistic mechanics and Lorentz transformations, which were first discovered by Hendrik Lorentz, F=\gamma ma (\gamma is the Lorentz factor, which is almost equal to 1 for low speeds).

General relativistic versus quantum

Relativistic corrections are also needed for quantum mechanics, although general relativity has not been integrated. The two theories remain incompatible, a hurdle which must be overcome in developing a theory of everything.

History

Antiquity

The main theory of mechanics in antiquity was Aristotelian mechanics.[4] A later developer in this tradition is Hipparchus.[5]

Medieval age


Arabic Machine Manuscript. Unknown date (at a guess: 16th to 19th centuries).
In the Middle Ages, Aristotle's theories were criticized and modified by a number of figures, beginning with John Philoponus in the 6th century. A central problem was that of projectile motion, which was discussed by Hipparchus and Philoponus. This led to the development of the theory of impetus by 14th century French Jean Buridan, which developed into the modern theories of inertia, velocity, acceleration and momentum. This work and others was developed in 14th century England by the Oxford Calculators such as Thomas Bradwardine, who studied and formulated various laws regarding falling bodies.
On the question of a body subject to a constant (uniform) force, the 12th century Jewish-Arab Nathanel (Iraqi, of Baghdad) stated that constant force imparts constant acceleration, while the main properties are uniformly accelerated motion (as of falling bodies) was worked out by the 14th century Oxford Calculators.

Early modern age

Two central figures in the early modern age are Galileo Galilei and Isaac Newton. Galileo's final statement of his mechanics, particularly of falling bodies, is his Two New Sciences (1638). Newton's 1687 Philosophiæ Naturalis Principia Mathematica provided a detailed mathematical account of mechanics, using the newly developed mathematics of calculus and providing the basis of Newtonian mechanics.[5]
There is some dispute over priority of various ideas: Newton's Principia is certainly the seminal work and has been tremendously influential, and the systematic mathematics therein did not and could not have been stated earlier because calculus had not been developed. However, many of the ideas, particularly as pertain to inertia (impetus) and falling bodies had been developed and stated by earlier researchers, both the then-recent Galileo and the less-known medieval predecessors. Precise credit is at times difficult or contentious because scientific language and standards of proof changed, so whether medieval statements are equivalent to modern statements or sufficient proof, or instead similar to modern statements and hypotheses is often debatable.

Modern age

Two main modern developments in mechanics are general relativity of Einstein, and quantum mechanics, both developed in the 20th century based in part on earlier 19th century ideas.

Types of mechanical bodies

Thus the often-used term body needs to stand for a wide assortment of objects, including particles, projectiles, spacecraft, stars, parts of machinery, parts of solids, parts of fluids (gases and liquids), etc.
Other distinctions between the various sub-disciplines of mechanics, concern the nature of the bodies being described. Particles are bodies with little (known) internal structure, treated as mathematical points in classical mechanics. Rigid bodies have size and shape, but retain a simplicity close to that of the particle, adding just a few so-called degrees of freedom, such as orientation in space.
Otherwise, bodies may be semi-rigid, i.e. elastic, or non-rigid, i.e. fluid. These subjects have both classical and quantum divisions of study.
For instance, the motion of a spacecraft, regarding its orbit and attitude (rotation), is described by the relativistic theory of classical mechanics, while the analogous movements of an atomic nucleus are described by quantum mechanics.

Sub-disciplines in mechanics

The following are two lists of various subjects that are studied in mechanics.
Note that there is also the "theory of fields" which constitutes a separate discipline in physics, formally treated as distinct from mechanics, whether classical fields or quantum fields. But in actual practice, subjects belonging to mechanics and fields are closely interwoven. Thus, for instance, forces that act on particles are frequently derived from fields (electromagnetic or gravitational), and particles generate fields by acting as sources. In fact, in quantum mechanics, particles themselves are fields, as described theoretically by the wave function.

Classical mechanics

The following are described as forming classical mechanics:

Quantum mechanics

The following are categorized as being part of quantum mechanics:

Professional organizations

  • Applied Mechanics Division, American Society of Mechanical Engineers
  • Fluid Dynamics Division, American Physical Society
  • Institution of Mechanical Engineers is the United Kingdom's qualifying body for Mechanical Engineers and has been the home of Mechanical Engineers for over 150 years.
  • International Union of Theoretical and Applied Mechanics

    Optics

    From Wikipedia, the free encyclopedia

    Optics includes study of dispersion of light.
    Optics is the branch of physics which involves the behaviour and properties of light, including its interactions with matter and the construction of instruments that use or detect it.[1] Optics usually describes the behaviour of visible, ultraviolet, and infrared light. Because light is an electromagnetic wave, other forms of electromagnetic radiation such as X-rays, microwaves, and radio waves exhibit similar properties.[1]
    Most optical phenomena can be accounted for using the classical electromagnetic description of light. Complete electromagnetic descriptions of light are, however, often difficult to apply in practice. Practical optics is usually done using simplified models. The most common of these, geometric optics, treats light as a collection of rays that travel in straight lines and bend when they pass through or reflect from surfaces. Physical optics is a more comprehensive model of light, which includes wave effects such as diffraction and interference that cannot be accounted for in geometric optics. Historically, the ray-based model of light was developed first, followed by the wave model of light. Progress in electromagnetic theory in the 19th century led to the discovery that light waves were in fact electromagnetic radiation.
    Some phenomena depend on the fact that light has both wave-like and particle-like properties. Explanation of these effects requires quantum mechanics. When considering light's particle-like properties, the light is modelled as a collection of particles called "photons". Quantum optics deals with the application of quantum mechanics to optical systems.
    Optical science is relevant to and studied in many related disciplines including astronomy, various engineering fields, photography, and medicine (particularly ophthalmology and optometry). Practical applications of optics are found in a variety of technologies and everyday objects, including mirrors, lenses, telescopes, microscopes, lasers, and fibre optics.

    History


    The Nimrud lens
    Optics began with the development of lenses by the ancient Egyptians and Mesopotamians. The earliest known lenses, made from polished crystal, often quartz, date from as early as 700 BC for Assyrian lenses such as the Layard/Nimrud lens.[2] The ancient Romans and Greeks filled glass spheres with water to make lenses. These practical developments were followed by the development of theories of light and vision by ancient Greek and Indian philosophers, and the development of geometrical optics in the Greco-Roman world. The word optics comes from the ancient Greek word ὀπτική, meaning "appearance, look".[3]
    Greek philosophy on optics broke down into two opposing theories on how vision worked, the "intro-mission theory" and the "emission theory".[4] The intro-mission approach saw vision as coming from objects casting off copies of themselves (called eidola) that were captured by the eye. With many propagators including Democritus, Epicurus, Aristotle and their followers, this theory seems to have some contact with modern theories of what vision really is, but it remained only speculation lacking any experimental foundation.
    Plato first articulated the emission theory, the idea that visual perception is accomplished by rays emitted by the eyes. He also commented on the parity reversal of mirrors in Timaeus.[5] Some hundred years later, Euclid wrote a treatise entitled Optics where he linked vision to geometry, creating geometrical optics.[6] He based his work on Plato's emission theory wherein he described the mathematical rules of perspective and described the effects of refraction qualitatively, although he questioned that a beam of light from the eye could instantaneously light up the stars every time someone blinked.[7] Ptolemy, in his treatise Optics, held an extramission-intromission theory of vision: the rays (or flux) from the eye formed a cone, the vertex being within the eye, and the base defining the visual field. The rays were sensitive, and conveyed information back to the observer’s intellect about the distance and orientation of surfaces. He summarised much of Euclid and went on to describe a way to measure the angle of refraction, though he failed to notice the empirical relationship between it and the angle of incidence.[8]

    Reproduction of a page of Ibn Sahl's manuscript showing his knowledge of the law of refraction, now known as Snell's law
    During the Middle Ages, Greek ideas about optics were resurrected and extended by writers in the Muslim world. One of the earliest of these was Al-Kindi (c. 801–73) who wrote on the merits of Aristotelian and Euclidean ideas of optics, favouring the emission theory since it could better quantify optical phenomenon.[9] In 984, the Persian mathematician Ibn Sahl wrote the treatise "On burning mirrors and lenses", correctly describing a law of refraction equivalent to Snell's law.[10] He used this law to compute optimum shapes for lenses and curved mirrors. In the early 11th century, Alhazen (Ibn al-Haytham) wrote the Book of Optics (Kitab al-manazir) in which he explored reflection and refraction and proposed a new system for explaining vision and light based on observation and experiment.[11][12][13][14][15] He rejected the "emission theory" of Ptolemaic optics with its rays being emitted by the eye, and instead put forward the idea that light reflected in all directions in straight lines from all points of the objects being viewed and then entered the eye, although he was unable to correctly explain how the eye captured the rays.[16] Alhazen's work was largely ignored in the Arabic world but it was anonymously translated into Latin around 1200 A.D. and further summarised and expanded on by the Polish monk Witelo[17] making it a standard text on optics in Europe for the next 400 years.
    In the 13th century medieval Europe the English bishop Robert Grosseteste wrote on a wide range of scientific topics discussing light from four different perspectives: an epistemology of light, a metaphysics or cosmogony of light, an etiology or physics of light, and a theology of light,[18] basing it on the works Aristotle and Platonism. Grosseteste's most famous disciple, Roger Bacon, wrote works citing a wide range of recently translated optical and philosophical works, including those of Alhazen, Aristotle, Avicenna, Averroes, Euclid, al-Kindi, Ptolemy, Tideus, and Constantine the African. Bacon was able to use parts of glass spheres as magnifying glasses to demonstrate that light reflects from objects rather than being released from them.
    In Italy, around 1284, Salvino D'Armate invented the first wearable eyeglasses.[19] This was the start of the optical industry of grinding and polishing lenses for these "spectacles", first in Venice and Florence in the thirteenth century,[20] and later in the spectacle making centres in both the Netherlands and Germany.[21] Spectacle makers created improved types of lenses for the correction of vision based more on empirical knowledge gained from observing the effects of the lenses rather than using the rudimentary optical theory of the day (theory which for the most part could not even adequately explain how spectacles worked).[22][23] This practical development, mastery, and experimentation with lenses led directly to the invention of the compound optical microscope around 1595, and the refracting telescope in 1608, both of which appeared in the spectacle making centres in the Netherlands.[24][25]
    In the early 17th century Johannes Kepler expanded on geometric optics in his writings, covering lenses, reflection by flat and curved mirrors, the principles of pinhole cameras, inverse-square law governing the intensity of light, and the optical explanations of astronomical phenomena such as lunar and solar eclipses and astronomical parallax. He was also able to correctly deduce the role of the retina as the actual organ that recorded images, finally being able to scientifically quantify the effects of different types of lenses that spectacle makers had been observing over the previous 300 years.[26] After the invention of the telescope Kepler set out the theoretical basis on how they worked and described an improved version, known as the Keplerian telescope, using two convex lenses to produce higher magnification.[27]

    Cover of the first edition of Newton's Opticks
    Optical theory progressed in the mid-17th century with treatises written by philosopher René Descartes, which explained a variety of optical phenomena including reflection and refraction by assuming that light was emitted by objects which produced it.[28] This differed substantively from the ancient Greek emission theory. In the late 1660s and early 1670s, Newton expanded Descartes' ideas into a corpuscle theory of light, famously determining that white light was a mix of colours which can be separated into its component parts with a prism. In 1690, Christiaan Huygens proposed a wave theory for light based on suggestions that had been made by Robert Hooke in 1664. Hooke himself publicly criticised Newton's theories of light and the feud between the two lasted until Hooke's death. In 1704, Newton published Opticks and, at the time, partly because of his success in other areas of physics, he was generally considered to be the victor in the debate over the nature of light.[28]
    Newtonian optics was generally accepted until the early 19th century when Thomas Young and Augustin-Jean Fresnel conducted experiments on the interference of light that firmly established light's wave nature. Young's famous double slit experiment showed that light followed the law of superposition, which is a wave-like property not predicted by Newton's corpuscle theory. This work led to a theory of diffraction for light and opened an entire area of study in physical optics.[29] Wave optics was successfully unified with electromagnetic theory by James Clerk Maxwell in the 1860s.[30]
    The next development in optical theory came in 1899 when Max Planck correctly modelled blackbody radiation by assuming that the exchange of energy between light and matter only occurred in discrete amounts he called quanta.[31] In 1905 Albert Einstein published the theory of the photoelectric effect that firmly established the quantization of light itself.[32][33] In 1913 Niels Bohr showed that atoms could only emit discrete amounts of energy, thus explaining the discrete lines seen in emission and absorption spectra.[34] The understanding of the interaction between light and matter which followed from these developments not only formed the basis of quantum optics but also was crucial for the development of quantum mechanics as a whole. The ultimate culmination, the theory of quantum electrodynamics, explains all optics and electromagnetic processes in general as the result of the exchange of real and virtual photons.[35]
    Quantum optics gained practical importance with the inventions of the maser in 1953 and of the laser in 1960.[36] Following the work of Paul Dirac in quantum field theory, George Sudarshan, Roy J. Glauber, and Leonard Mandel applied quantum theory to the electromagnetic field in the 1950s and 1960s to gain a more detailed understanding of photodetection and the statistics of light.

    Classical optics

    Classical optics is divided into two main branches: geometrical optics and physical optics. In geometrical, or ray optics, light is considered to travel in straight lines, and in physical, or wave optics, light is considered to be an electromagnetic wave.
    Geometrical optics can be viewed as an approximation of physical optics which can be applied when the wavelength of the light used is much smaller than the size of the optical elements or system being modelled.

    Geometrical optics


    Geometry of reflection and refraction of light rays
    Geometrical optics, or ray optics, describes the propagation of light in terms of "rays" which travel in straight lines, and whose paths are governed by the laws of reflection and refraction at interfaces between different media.[37] These laws were discovered empirically as far back as 984 AD[10] and have been used in the design of optical components and instruments from then until the present day. They can be summarised as follows:
    When a ray of light hits the boundary between two transparent materials, it is divided into a reflected and a refracted ray.
    The law of reflection says that the reflected ray lies in the plane of incidence, and the angle of reflection equals the angle of incidence.
    The law of refraction says that the refracted ray lies in the plane of incidence, and the sine of the angle of refraction divided by the sine of the angle of incidence is a constant.
    \frac {\sin {\theta_1}}{\sin {\theta_2}} = n
    where n is a constant for any two materials and a given colour of light. It is known as the refractive index.
    The laws of reflection and refraction can be derived from Fermat's principle which states that the path taken between two points by a ray of light is the path that can be traversed in the least time.[38]

    Approximations

    Geometric optics is often simplified by making the paraxial approximation, or "small angle approximation." The mathematical behaviour then becomes linear, allowing optical components and systems to be described by simple matrices. This leads to the techniques of Gaussian optics and paraxial ray tracing, which are used to find basic properties of optical systems, such as approximate image and object positions and magnifications.[39]

    Reflections


    Diagram of specular reflection
    Reflections can be divided into two types: specular reflection and diffuse reflection. Specular reflection describes the gloss of surfaces such as mirrors, which reflect light in a simple, predictable way. This allows for production of reflected images that can be associated with an actual (real) or extrapolated (virtual) location in space. Diffuse reflection describes opaque, non limpid materials, such as paper or rock. The reflections from these surfaces can only be described statistically, with the exact distribution of the reflected light depending on the microscopic structure of the material. Many diffuse reflectors are described or can be approximated by Lambert's cosine law, which describes surfaces that have equal luminance when viewed from any angle. Glossy surfaces can give both specular and diffuse reflection.
    In specular reflection, the direction of the reflected ray is determined by the angle the incident ray makes with the surface normal, a line perpendicular to the surface at the point where the ray hits. The incident and reflected rays and the normal lie in a single plane, and the angle between the reflected ray and the surface normal is the same as that between the incident ray and the normal.[40] This is known as the Law of Reflection.
    For flat mirrors, the law of reflection implies that images of objects are upright and the same distance behind the mirror as the objects are in front of the mirror. The image size is the same as the object size. The law also implies that mirror images are parity inverted, which we perceive as a left-right inversion. Images formed from reflection in two (or any even number of) mirrors are not parity inverted. Corner reflectors[40] retroreflect light, producing reflected rays that travel back in the direction from which the incident rays came.
    Mirrors with curved surfaces can be modelled by ray-tracing and using the law of reflection at each point on the surface. For mirrors with parabolic surfaces, parallel rays incident on the mirror produce reflected rays that converge at a common focus. Other curved surfaces may also focus light, but with aberrations due to the diverging shape causing the focus to be smeared out in space. In particular, spherical mirrors exhibit spherical aberration. Curved mirrors can form images with magnification greater than or less than one, and the magnification can be negative, indicating that the image is inverted. An upright image formed by reflection in a mirror is always virtual, while an inverted image is real and can be projected onto a screen.[40]

    Refractions


    Illustration of Snell's Law for the case n1 < n2, such as air/water interface
    Refraction occurs when light travels through an area of space that has a changing index of refraction; this principle allows for lenses and the focusing of light. The simplest case of refraction occurs when there is an interface between a uniform medium with index of refraction n_1 and another medium with index of refraction n_2. In such situations, Snell's Law describes the resulting deflection of the light ray:
    n_1\sin\theta_1 = n_2\sin\theta_2\
    where \theta_1 and \theta_2 are the angles between the normal (to the interface) and the incident and refracted waves, respectively. This phenomenon is also associated with a changing speed of light as seen from the definition of index of refraction provided above which implies:
    v_1\sin\theta_2\ = v_2\sin\theta_1
    where v_1 and v_2 are the wave velocities through the respective media.[40]
    Various consequences of Snell's Law include the fact that for light rays travelling from a material with a high index of refraction to a material with a low index of refraction, it is possible for the interaction with the interface to result in zero transmission. This phenomenon is called total internal reflection and allows for fibre optics technology. As light signals travel down a fibre optic cable, it undergoes total internal reflection allowing for essentially no light lost over the length of the cable. It is also possible to produce polarised light rays using a combination of reflection and refraction: When a refracted ray and the reflected ray form a right angle, the reflected ray has the property of "plane polarization". The angle of incidence required for such a scenario is known as Brewster's angle.[40]
    Snell's Law can be used to predict the deflection of light rays as they pass through "linear media" as long as the indexes of refraction and the geometry of the media are known. For example, the propagation of light through a prism results in the light ray being deflected depending on the shape and orientation of the prism. Additionally, since different frequencies of light have slightly different indexes of refraction in most materials, refraction can be used to produce dispersion spectra that appear as rainbows. The discovery of this phenomenon when passing light through a prism is famously attributed to Isaac Newton.[40]
    Some media have an index of refraction which varies gradually with position and, thus, light rays curve through the medium rather than travel in straight lines. This effect is what is responsible for mirages seen on hot days where the changing index of refraction of the air causes the light rays to bend creating the appearance of specular reflections in the distance (as if on the surface of a pool of water). Material that has a varying index of refraction is called a gradient-index (GRIN) material and has many useful properties used in modern optical scanning technologies including photocopiers and scanners. The phenomenon is studied in the field of gradient-index optics.[41]

    A ray tracing diagram for a converging lens.
    A device which produces converging or diverging light rays due to refraction is known as a lens. Thin lenses produce focal points on either side that can be modelled using the lensmaker's equation.[42] In general, two types of lenses exist: convex lenses, which cause parallel light rays to converge, and concave lenses, which cause parallel light rays to diverge. The detailed prediction of how images are produced by these lenses can be made using ray-tracing similar to curved mirrors. Similarly to curved mirrors, thin lenses follow a simple equation that determines the location of the images given a particular focal length (f) and object distance (S_1):
    \frac{1}{S_1} + \frac{1}{S_2} = \frac{1}{f}
    where S_2 is the distance associated with the image and is considered by convention to be negative if on the same side of the lens as the object and positive if on the opposite side of the lens.[42] The focal length f is considered negative for concave lenses.
    Lens1.svg
    Incoming parallel rays are focused by a convex lens into an inverted real image one focal length from the lens, on the far side of the lens. Rays from an object at finite distance are focused further from the lens than the focal distance; the closer the object is to the lens, the further the image is from the lens. With concave lenses, incoming parallel rays diverge after going through the lens, in such a way that they seem to have originated at an upright virtual image one focal length from the lens, on the same side of the lens that the parallel rays are approaching on. Rays from an object at finite distance are associated with a virtual image that is closer to the lens than the focal length, and on the same side of the lens as the object. The closer the object is to the lens, the closer the virtual image is to the lens.
    Likewise, the magnification of a lens is given by
     M = - \frac{S_2}{S_1} = \frac{f}{f - S_1}
    where the negative sign is given, by convention, to indicate an upright object for positive values and an inverted object for negative values. Similar to mirrors, upright images produced by single lenses are virtual while inverted images are real.[40]
    Lenses suffer from aberrations that distort images and focal points. These are due to both to geometrical imperfections and due to the changing index of refraction for different wavelengths of light (chromatic aberration).[40]

    Images of black letters in a thin convex lens of focal length f  are shown in red. Selected rays are shown for letters E, I and K in blue, green and orange, respectively. Note that E (at 2f) has an equal-size, real and inverted image; I (at f) has its image at infinity; and K (at f/2) has a double-size, virtual and upright image.

    Physical optics

    In physical optics, light is considered to propagate as a wave. This model predicts phenomena such as interference and diffraction, which are not explained by geometric optics. The speed of light waves in air is approximately 3.0×108 m/s (exactly 299,792,458 m/s in vacuum). The wavelength of visible light waves varies between 400 and 700 nm, but the term "light" is also often applied to infrared (0.7–300 μm) and ultraviolet radiation (10–400 nm).
    The wave model can be used to make predictions about how an optical system will behave without requiring an explanation of what is "waving" in what medium. Until the middle of the 19th century, most physicists believed in an "ethereal" medium in which the light disturbance propagated.[43] The existence of electromagnetic waves was predicted in 1865 by Maxwell's equations. These waves propagate at the speed of light and have varying electric and magnetic fields which are orthogonal to one another, and also to the direction of propagation of the waves.[44] Light waves are now generally treated as electromagnetic waves except when quantum mechanical effects have to be considered.

    Modelling and design of optical systems using physical optics

    Many simplified approximations are available for analysing and designing optical systems. Most of these use a single scalar quantity to represent the electric field of the light wave, rather than using a vector model with orthogonal electric and magnetic vectors.[45] The Huygens–Fresnel equation is one such model. This was derived empirically by Fresnel in 1815, based on Huygen's hypothesis that each point on a wavefront generates a secondary spherical wavefront, which Fresnel combined with the principle of superposition of waves. The Kirchhoff diffraction equation, which is derived using Maxwell's equations, puts the Huygens-Fresnel equation on a firmer physical foundation. Examples of the application of Huygens–Fresnel principle can be found in the sections on diffraction and Fraunhofer diffraction.
    More rigorous models, involving the modelling of both electric and magnetic fields of the light wave, are required when dealing with the detailed interaction of light with materials where the interaction depends on their electric and magnetic properties. For instance, the behaviour of a light wave interacting with a metal surface is quite different from what happens when it interacts with a dielectric material. A vector model must also be used to model polarised light.
    Numerical modeling techniques such as the finite element method, the boundary element method and the transmission-line matrix method can be used to model the propagation of light in systems which cannot be solved analytically. Such models are computationally demanding and are normally only used to solve small-scale problems that require accuracy beyond that which can be achieved with analytical solutions.[46]
    All of the results from geometrical optics can be recovered using the techniques of Fourier optics which apply many of the same mathematical and analytical techniques used in acoustic engineering and signal processing.
    Gaussian beam propagation is a simple paraxial physical optics model for the propagation of coherent radiation such as laser beams. This technique partially accounts for diffraction, allowing accurate calculations of the rate at which a laser beam expands with distance, and the minimum size to which the beam can be focused. Gaussian beam propagation thus bridges the gap between geometric and physical optics.[47]

    Superposition and interference

    In the absence of nonlinear effects, the superposition principle can be used to predict the shape of interacting waveforms through the simple addition of the disturbances.[48] This interaction of waves to produce a resulting pattern is generally termed "interference" and can result in a variety of outcomes. If two waves of the same wavelength and frequency are in phase, both the wave crests and wave troughs align. This results in constructive interference and an increase in the amplitude of the wave, which for light is associated with a brightening of the waveform in that location. Alternatively, if the two waves of the same wavelength and frequency are out of phase, then the wave crests will align with wave troughs and vice-versa. This results in destructive interference and a decrease in the amplitude of the wave, which for light is associated with a dimming of the waveform at that location. See below for an illustration of this effect.[48]
    combined
    waveform
    Interference of two waves.svg
    wave 1
    wave 2

    Two waves in phase Two waves 180° out
    of phase

    When oil or fuel is spilled, colourful patterns are formed by thin-film interference.
    Since the Huygens–Fresnel principle states that every point of a wavefront is associated with the production of a new disturbance, it is possible for a wavefront to interfere with itself constructively or destructively at different locations producing bright and dark fringes in regular and predictable patterns.[48] Interferometry is the science of measuring these patterns, usually as a means of making precise determinations of distances or angular resolutions.[49] The Michelson interferometer was a famous instrument which used interference effects to accurately measure the speed of light.[50]
    The appearance of thin films and coatings is directly affected by interference effects. Antireflective coatings use destructive interference to reduce the reflectivity of the surfaces they coat, and can be used to minimise glare and unwanted reflections. The simplest case is a single layer with thickness one-fourth the wavelength of incident light. The reflected wave from the top of the film and the reflected wave from the film/material interface are then exactly 180° out of phase, causing destructive interference. The waves are only exactly out of phase for one wavelength, which would typically be chosen to be near the centre of the visible spectrum, around 550 nm. More complex designs using multiple layers can achieve low reflectivity over a broad band, or extremely low reflectivity at a single wavelength.
    Constructive interference in thin films can create strong reflection of light in a range of wavelengths, which can be narrow or broad depending on the design of the coating. These films are used to make dielectric mirrors, interference filters, heat reflectors, and filters for colour separation in colour television cameras. This interference effect is also what causes the colourful rainbow patterns seen in oil slicks.[48]

    Diffraction and optical resolution


    Diffraction on two slits separated by distance d. The bright fringes occur along lines where black lines intersect with black lines and white lines intersect with white lines. These fringes are separated by angle \theta and are numbered as order n.
    Diffraction is the process by which light interference is most commonly observed. The effect was first described in 1665 by Francesco Maria Grimaldi, who also coined the term from the Latin diffringere, 'to break into pieces'.[51][52] Later that century, Robert Hooke and Isaac Newton also described phenomena now known to be diffraction in Newton's rings[53] while James Gregory recorded his observations of diffraction patterns from bird feathers.[54]
    The first physical optics model of diffraction that relied on the Huygens–Fresnel principle was developed in 1803 by Thomas Young in his interference experiments with the interference patterns of two closely spaced slits. Young showed that his results could only be explained if the two slits acted as two unique sources of waves rather than corpuscles.[55] In 1815 and 1818, Augustin-Jean Fresnel firmly established the mathematics of how wave interference can account for diffraction.[42]
    The simplest physical models of diffraction use equations that describe the angular separation of light and dark fringes due to light of a particular wavelength (λ). In general, the equation takes the form
    m \lambda = d \sin \theta
    where d is the separation between two wavefront sources (in the case of Young's experiments, it was two slits), \theta is the angular separation between the central fringe and the mth order fringe, where the central maximum is m = 0.[56]
    This equation is modified slightly to take into account a variety of situations such as diffraction through a single gap, diffraction through multiple slits, or diffraction through a diffraction grating that contains a large number of slits at equal spacing.[56] More complicated models of diffraction require working with the mathematics of Fresnel or Fraunhofer diffraction.[57]
    X-ray diffraction makes use of the fact that atoms in a crystal have regular spacing at distances that are on the order of one angstrom. To see diffraction patterns, x-rays with similar wavelengths to that spacing are passed through the crystal. Since crystals are three-dimensional objects rather than two-dimensional gratings, the associated diffraction pattern varies in two directions according to Bragg reflection, with the associated bright spots occurring in unique patterns and d being twice the spacing between atoms.[56]
    Diffraction effects limit the ability for an optical detector to optically resolve separate light sources. In general, light that is passing through an aperture will experience diffraction and the best images that can be created (as described in diffraction-limited optics) appear as a central spot with surrounding bright rings, separated by dark nulls; this pattern is known as an Airy pattern, and the central bright lobe as an Airy disk.[42] The size of such a disk is given by
     \sin \theta = 1.22 \frac{\lambda}{D}
    where θ is the angular resolution, λ is the wavelength of the light, and D is the diameter of the lens aperture. If the angular separation of the two points is significantly less than the Airy disk angular radius, then the two points cannot be resolved in the image, but if their angular separation is much greater than this, distinct images of the two points are formed and they can therefore be resolved. Rayleigh defined the somewhat arbitrary "Rayleigh criterion" that two points whose angular separation is equal to the Airy disk radius (measured to first null, that is, to the first place where no light is seen) can be considered to be resolved. It can be seen that the greater the diameter of the lens or its aperture, the finer the resolution.[56] Interferometry, with its ability to mimic extremely large baseline apertures, allows for the greatest angular resolution possible.[49]
    For astronomical imaging, the atmosphere prevents optimal resolution from being achieved in the visible spectrum due to the atmospheric scattering and dispersion which cause stars to twinkle. Astronomers refer to this effect as the quality of astronomical seeing. Techniques known as adaptive optics have been used to eliminate the atmospheric disruption of images and achieve results that approach the diffraction limit.[58]

    Dispersion and scattering


    Conceptual animation of light dispersion through a prism. High frequency (blue) light is deflected the most, and low frequency (red) the least.
    Refractive processes take place in the physical optics limit, where the wavelength of light is similar to other distances, as a kind of scattering. The simplest type of scattering is Thomson scattering which occurs when electromagnetic waves are deflected by single particles. In the limit of Thompson scattering, in which the wavelike nature of light is evident, light is dispersed independent of the frequency, in contrast to Compton scattering which is frequency-dependent and strictly a quantum mechanical process, involving the nature of light as particles. In a statistical sense, elastic scattering of light by numerous particles much smaller than the wavelength of the light is a process known as Rayleigh scattering while the similar process for scattering by particles that are similar or larger in wavelength is known as Mie scattering with the Tyndall effect being a commonly observed result. A small proportion of light scattering from atoms or molecules may undergo Raman scattering, wherein the frequency changes due to excitation of the atoms and molecules. Brillouin scattering occurs when the frequency of light changes due to local changes with time and movements of a dense material.[59]
    Dispersion occurs when different frequencies of light have different phase velocities, due either to material properties (material dispersion) or to the geometry of an optical waveguide (waveguide dispersion). The most familiar form of dispersion is a decrease in index of refraction with increasing wavelength, which is seen in most transparent materials. This is called "normal dispersion". It occurs in all dielectric materials, in wavelength ranges where the material does not absorb light.[60] In wavelength ranges where a medium has significant absorption, the index of refraction can increase with wavelength. This is called "anomalous dispersion".[40][60]
    The separation of colours by a prism is an example of normal dispersion. At the surfaces of the prism, Snell's law predicts that light incident at an angle θ to the normal will be refracted at an angle arcsin(sin (θ) / n). Thus, blue light, with its higher refractive index, is bent more strongly than red light, resulting in the well-known rainbow pattern.[40]

    Dispersion: two sinusoids propagating at different speeds make a moving interference pattern. The red dot moves with the phase velocity, and the green dots propagate with the group velocity. In this case, the phase velocity is twice the group velocity. The red dot overtakes two green dots, when moving from the left to the right of the figure. In effect, the individual waves (which travel with the phase velocity) escape from the wave packet (which travels with the group velocity).
    Material dispersion is often characterised by the Abbe number, which gives a simple measure of dispersion based on the index of refraction at three specific wavelengths. Waveguide dispersion is dependent on the propagation constant.[42] Both kinds of dispersion cause changes in the group characteristics of the wave, the features of the wave packet that change with the same frequency as the amplitude of the electromagnetic wave. "Group velocity dispersion" manifests as a spreading-out of the signal "envelope" of the radiation and can be quantified with a group dispersion delay parameter:
    D = \frac{1}{v_g^2} \frac{dv_g}{d\lambda}
    where v_g is the group velocity.[61] For a uniform medium, the group velocity is
    v_g = c \left( n - \lambda \frac{dn}{d\lambda} \right)^{-1}
    where n is the index of refraction and c is the speed of light in a vacuum.[62] This gives a simpler form for the dispersion delay parameter:
    D = - \frac{\lambda}{c} \, \frac{d^2 n}{d \lambda^2}.
    If D is less than zero, the medium is said to have positive dispersion or normal dispersion. If D is greater than zero, the medium has negative dispersion. If a light pulse is propagated through a normally dispersive medium, the result is the higher frequency components slow down more than the lower frequency components. The pulse therefore becomes positively chirped, or up-chirped, increasing in frequency with time. This causes the spectrum coming out of a prism to appear with red light the least refracted and blue/violet light the most refracted. Conversely, if a pulse travels through an anomalously (negatively) dispersive medium, high frequency components travel faster than the lower ones, and the pulse becomes negatively chirped, or down-chirped, decreasing in frequency with time.[63]
    The result of group velocity dispersion, whether negative or positive, is ultimately temporal spreading of the pulse. This makes dispersion management extremely important in optical communications systems based on optical fibres, since if dispersion is too high, a group of pulses representing information will each spread in time and merge, making it impossible to extract the signal.[61]

    Polarization

    Polarization is a general property of waves that describes the orientation of their oscillations. For transverse waves such as many electromagnetic waves, it describes the orientation of the oscillations in the plane perpendicular to the wave's direction of travel. The oscillations may be oriented in a single direction (linear polarization), or the oscillation direction may rotate as the wave travels (circular or elliptical polarization). Circularly polarised waves can rotate rightward or leftward in the direction of travel, and which of those two rotations is present in a wave is called the wave's chirality.[64]
    The typical way to consider polarization is to keep track of the orientation of the electric field vector as the electromagnetic wave propagates. The electric field vector of a plane wave may be arbitrarily divided into two perpendicular components labeled x and y (with z indicating the direction of travel). The shape traced out in the x-y plane by the electric field vector is a Lissajous figure that describes the polarization state.[42] The following figures show some examples of the evolution of the electric field vector (blue), with time (the vertical axes), at a particular point in space, along with its x and y components (red/left and green/right), and the path traced by the vector in the plane (purple): The same evolution would occur when looking at the electric field at a particular time while evolving the point in space, along the direction opposite to propagation.
    Linear polarization diagram
    Linear
    Circular polarization diagram
    Circular
    Elliptical polarization diagram
    Elliptical polarization

    In the leftmost figure above, the x and y components of the light wave are in phase. In this case, the ratio of their strengths is constant, so the direction of the electric vector (the vector sum of these two components) is constant. Since the tip of the vector traces out a single line in the plane, this special case is called linear polarization. The direction of this line depends on the relative amplitudes of the two components.[64]
    In the middle figure, the two orthogonal components have the same amplitudes and are 90° out of phase. In this case, one component is zero when the other component is at maximum or minimum amplitude. There are two possible phase relationships that satisfy this requirement: the x component can be 90° ahead of the y component or it can be 90° behind the y component. In this special case, the electric vector traces out a circle in the plane, so this polarization is called circular polarization. The rotation direction in the circle depends on which of the two phase relationships exists and corresponds to right-hand circular polarization and left-hand circular polarization.[42]
    In all other cases, where the two components either do not have the same amplitudes and/or their phase difference is neither zero nor a multiple of 90°, the polarization is called elliptical polarization because the electric vector traces out an ellipse in the plane (the polarization ellipse). This is shown in the above figure on the right. Detailed mathematics of polarization is done using Jones calculus and is characterised by the Stokes parameters.[42]
    Changing polarization
    Media that have different indexes of refraction for different polarization modes are called birefringent.[64] Well known manifestations of this effect appear in optical wave plates/retarders (linear modes) and in Faraday rotation/optical rotation (circular modes).[42] If the path length in the birefringent medium is sufficient, plane waves will exit the material with a significantly different propagation direction, due to refraction. For example, this is the case with macroscopic crystals of calcite, which present the viewer with two offset, orthogonally polarised images of whatever is viewed through them. It was this effect that provided the first discovery of polarization, by Erasmus Bartholinus in 1669. In addition, the phase shift, and thus the change in polarization state, is usually frequency dependent, which, in combination with dichroism, often gives rise to bright colours and rainbow-like effects. In mineralogy, such properties, known as pleochroism, are frequently exploited for the purpose of identifying minerals using polarization microscopes. Additionally, many plastics that are not normally birefringent will become so when subject to mechanical stress, a phenomenon which is the basis of photoelasticity.[64] Non-birefringent methods, to rotate the linear polarization of light beams, include the use of prismatic polarization rotators which use total internal reflection in a prism set designed for efficient collinear transmission.[65]

    A polariser changing the orientation of linearly polarised light.
    In this picture, θ1θ0 = θi.
    Media that reduce the amplitude of certain polarization modes are called dichroic. with devices that block nearly all of the radiation in one mode known as polarizing filters or simply "polarisers". Malus' law, which is named after Étienne-Louis Malus, says that when a perfect polariser is placed in a linear polarised beam of light, the intensity, I, of the light that passes through is given by
     I = I_0 \cos^2 \theta_i \quad ,
    where
    I0 is the initial intensity,
    and θi is the angle between the light's initial polarization direction and the axis of the polariser.[64]
    A beam of unpolarised light can be thought of as containing a uniform mixture of linear polarizations at all possible angles. Since the average value of \cos^2 \theta is 1/2, the transmission coefficient becomes
     \frac {I}{I_0} = \frac {1}{2}\quad
    In practice, some light is lost in the polariser and the actual transmission of unpolarised light will be somewhat lower than this, around 38% for Polaroid-type polarisers but considerably higher (>49.9%) for some birefringent prism types.[42]
    In addition to birefringence and dichroism in extended media, polarization effects can also occur at the (reflective) interface between two materials of different refractive index. These effects are treated by the Fresnel equations. Part of the wave is transmitted and part is reflected, with the ratio depending on angle of incidence and the angle of refraction. In this way, physical optics recovers Brewster's angle.[42] When light reflects from a thin film on a surface, interference between the reflections from the film's surfaces can produce polarization in the reflected and transmitted light.
    Natural light

    The effects of a polarising filter on the sky in a photograph. Left picture is taken without polariser. For the right picture, filter was adjusted to eliminate certain polarizations of the scattered blue light from the sky.
    Most sources of electromagnetic radiation contain a large number of atoms or molecules that emit light. The orientation of the electric fields produced by these emitters may not be correlated, in which case the light is said to be unpolarised. If there is partial correlation between the emitters, the light is partially polarised. If the polarization is consistent across the spectrum of the source, partially polarised light can be described as a superposition of a completely unpolarised component, and a completely polarised one. One may then describe the light in terms of the degree of polarization, and the parameters of the polarization ellipse.[42]
    Light reflected by shiny transparent materials is partly or fully polarised, except when the light is normal (perpendicular) to the surface. It was this effect that allowed the mathematician Étienne-Louis Malus to make the measurements that allowed for his development of the first mathematical models for polarised light. Polarization occurs when light is scattered in the atmosphere. The scattered light produces the brightness and colour in clear skies. This partial polarization of scattered light can be taken advantage of using polarizing filters to darken the sky in photographs. Optical polarization is principally of importance in chemistry due to circular dichroism and optical rotation ("circular birefringence") exhibited by optically active (chiral) molecules.[42]

    Modern optics

    Modern optics encompasses the areas of optical science and engineering that became popular in the 20th century. These areas of optical science typically relate to the electromagnetic or quantum properties of light but do include other topics. A major subfield of modern optics, quantum optics, deals with specifically quantum mechanical properties of light. Quantum optics is not just theoretical; some modern devices, such as lasers, have principles of operation that depend on quantum mechanics. Light detectors, such as photomultipliers and channeltrons, respond to individual photons. Electronic image sensors, such as CCDs, exhibit shot noise corresponding to the statistics of individual photon events. Light-emitting diodes and photovoltaic cells, too, cannot be understood without quantum mechanics. In the study of these devices, quantum optics often overlaps with quantum electronics.[66]
    Specialty areas of optics research include the study of how light interacts with specific materials as in crystal optics and metamaterials. Other research focuses on the phenomenology of electromagnetic waves as in singular optics, non-imaging optics, non-linear optics, statistical optics, and radiometry. Additionally, computer engineers have taken an interest in integrated optics, machine vision, and photonic computing as possible components of the "next generation" of computers.[67]
    Today, the pure science of optics is called optical science or optical physics to distinguish it from applied optical sciences, which are referred to as optical engineering. Prominent subfields of optical engineering include illumination engineering, photonics, and optoelectronics with practical applications like lens design, fabrication and testing of optical components, and image processing. Some of these fields overlap, with nebulous boundaries between the subjects terms that mean slightly different things in different parts of the world and in different areas of industry. A professional community of researchers in nonlinear optics has developed in the last several decades due to advances in laser technology.[68]

    Lasers


    Experiments such as this one with high-power lasers are part of the modern optics research.
    A laser is a device that emits light (electromagnetic radiation) through a process called stimulated emission. The term laser is an acronym for Light Amplification by Stimulated Emission of Radiation.[69] Laser light is usually spatially coherent, which means that the light either is emitted in a narrow, low-divergence beam, or can be converted into one with the help of optical components such as lenses. Because the microwave equivalent of the laser, the maser, was developed first, devices that emit microwave and radio frequencies are usually called masers.[70]
    The first working laser was demonstrated on 16 May 1960 by Theodore Maiman at Hughes Research Laboratories.[71] When first invented, they were called "a solution looking for a problem".[72] Since then, lasers have become a multi-billion dollar industry, finding utility in thousands of highly varied applications. The first application of lasers visible in the daily lives of the general population was the supermarket barcode scanner, introduced in 1974.[73] The laserdisc player, introduced in 1978, was the first successful consumer product to include a laser, but the compact disc player was the first laser-equipped device to become truly common in consumers' homes, beginning in 1982.[74] These optical storage devices use a semiconductor laser less than a millimetre wide to scan the surface of the disc for data retrieval. Fibre-optic communication relies on lasers to transmit large amounts of information at the speed of light. Other common applications of lasers include laser printers and laser pointers. Lasers are used in medicine in areas such as bloodless surgery, laser eye surgery, and laser capture microdissection and in military applications such as missile defence systems, electro-optical countermeasures (EOCM), and LIDAR. Lasers are also used in holograms, bubblegrams, laser light shows, and laser hair removal.[75]

    Kapitsa–Dirac effect

    The Kapitsa–Dirac effect causes beams of particles to diffract as the result of meeting a standing wave of light. Light can be used to position matter using various phenomena (see optical tweezers).

    Applications

    Optics is part of everyday life. The ubiquity of visual systems in biology indicates the central role optics plays as the science of one of the five senses. Many people benefit from eyeglasses or contact lenses, and optics are integral to the functioning of many consumer goods including cameras. Rainbows and mirages are examples of optical phenomena. Optical communication provides the backbone for both the Internet and modern telephony.

    Human eye


    Model of a human eye. Features mentioned in this article are 3. ciliary muscle, 6. pupil, 8. cornea, 10. lens cortex, 22. optic nerve, 26. fovea, 30. retina
    The human eye functions by focusing light onto a layer of photoreceptor cells called the retina, which forms the inner lining of the back of the eye. The focusing is accomplished by a series of transparent media. Light entering the eye passes first through the cornea, which provides much of the eye's optical power. The light then continues through the fluid just behind the cornea—the anterior chamber, then passes through the pupil. The light then passes through the lens, which focuses the light further and allows adjustment of focus. The light then passes through the main body of fluid in the eye—the vitreous humour, and reaches the retina. The cells in the retina line the back of the eye, except for where the optic nerve exits; this results in a blind spot.
    There are two types of photoreceptor cells, rods and cones, which are sensitive to different aspects of light.[76] Rod cells are sensitive to the intensity of light over a wide frequency range, thus are responsible for black-and-white vision. Rod cells are not present on the fovea, the area of the retina responsible for central vision, and are not as responsive as cone cells to spatial and temporal changes in light. There are, however, twenty times more rod cells than cone cells in the retina because the rod cells are present across a wider area. Because of their wider distribution, rods are responsible for peripheral vision.[77]
    In contrast, cone cells are less sensitive to the overall intensity of light, but come in three varieties that are sensitive to different frequency-ranges and thus are used in the perception of colour and photopic vision. Cone cells are highly concentrated in the fovea and have a high visual acuity meaning that they are better at spatial resolution than rod cells. Since cone cells are not as sensitive to dim light as rod cells, most night vision is limited to rod cells. Likewise, since cone cells are in the fovea, central vision (including the vision needed to do most reading, fine detail work such as sewing, or careful examination of objects) is done by cone cells.[77]
    Ciliary muscles around the lens allow the eye's focus to be adjusted. This process is known as accommodation. The near point and far point define the nearest and farthest distances from the eye at which an object can be brought into sharp focus. For a person with normal vision, the far point is located at infinity. The near point's location depends on how much the muscles can increase the curvature of the lens, and how inflexible the lens has become with age. Optometrists, ophthalmologists, and opticians usually consider an appropriate near point to be closer than normal reading distance—approximately 25 cm.[76]
    Defects in vision can be explained using optical principles. As people age, the lens becomes less flexible and the near point recedes from the eye, a condition known as presbyopia. Similarly, people suffering from hyperopia cannot decrease the focal length of their lens enough to allow for nearby objects to be imaged on their retina. Conversely, people who cannot increase the focal length of their lens enough to allow for distant objects to be imaged on the retina suffer from myopia and have a far point that is considerably closer than infinity. A condition known as astigmatism results when the cornea is not spherical but instead is more curved in one direction. This causes horizontally extended objects to be focused on different parts of the retina than vertically extended objects, and results in distorted images.[76]
    All of these conditions can be corrected using corrective lenses. For presbyopia and hyperopia, a converging lens provides the extra curvature necessary to bring the near point closer to the eye while for myopia a diverging lens provides the curvature necessary to send the far point to infinity. Astigmatism is corrected with a cylindrical surface lens that curves more strongly in one direction than in another, compensating for the non-uniformity of the cornea.[78]
    The optical power of corrective lenses is measured in diopters, a value equal to the reciprocal of the focal length measured in meters; with a positive focal length corresponding to a converging lens and a negative focal length corresponding to a diverging lens. For lenses that correct for astigmatism as well, three numbers are given: one for the spherical power, one for the cylindrical power, and one for the angle of orientation of the astigmatism.[78]

    Visual effects


    The Ponzo Illusion relies on the fact that parallel lines appear to converge as they approach infinity.
    Optical illusions (also called visual illusions) are characterized by visually perceived images that differ from objective reality. The information gathered by the eye is processed in the brain to give a percept that differs from the object being imaged. Optical illusions can be the result of a variety of phenomena including physical effects that create images that are different from the objects that make them, the physiological effects on the eyes and brain of excessive stimulation (e.g. brightness, tilt, colour, movement), and cognitive illusions where the eye and brain make unconscious inferences.[79]
    Cognitive illusions include some which result from the unconscious misapplication of certain optical principles. For example, the Ames room, Hering, Müller-Lyer, Orbison, Ponzo, Sander, and Wundt illusions all rely on the suggestion of the appearance of distance by using converging and diverging lines, in the same way that parallel light rays (or indeed any set of parallel lines) appear to converge at a vanishing point at infinity in two-dimensionally rendered images with artistic perspective.[80] This suggestion is also responsible for the famous moon illusion where the moon, despite having essentially the same angular size, appears much larger near the horizon than it does at zenith.[81] This illusion so confounded Ptolemy that he incorrectly attributed it to atmospheric refraction when he described it in his treatise, Optics.[8]
    Another type of optical illusion exploits broken patterns to trick the mind into perceiving symmetries or asymmetries that are not present. Examples include the café wall, Ehrenstein, Fraser spiral, Poggendorff, and Zöllner illusions. Related, but not strictly illusions, are patterns that occur due to the superimposition of periodic structures. For example transparent tissues with a grid structure produce shapes known as moiré patterns, while the superimposition of periodic transparent patterns comprising parallel opaque lines or curves produces line moiré patterns.[82]

    Optical instruments


    Illustrations of various optical instruments from the 1728 Cyclopaedia
    Single lenses have a variety of applications including photographic lenses, corrective lenses, and magnifying glasses while single mirrors are used in parabolic reflectors and rear-view mirrors. Combining a number of mirrors, prisms, and lenses produces compound optical instruments which have practical uses. For example, a periscope is simply two plane mirrors aligned to allow for viewing around obstructions. The most famous compound optical instruments in science are the microscope and the telescope which were both invented by the Dutch in the late 16th century.[83]
    Microscopes were first developed with just two lenses: an objective lens and an eyepiece. The objective lens is essentially a magnifying glass and was designed with a very small focal length while the eyepiece generally has a longer focal length. This has the effect of producing magnified images of close objects. Generally, an additional source of illumination is used since magnified images are dimmer due to the conservation of energy and the spreading of light rays over a larger surface area. Modern microscopes, known as compound microscopes have many lenses in them (typically four) to optimize the functionality and enhance image stability.[83] A slightly different variety of microscope, the comparison microscope, looks at side-by-side images to produce a stereoscopic binocular view that appears three dimensional when used by humans.[84]
    The first telescopes, called refracting telescopes were also developed with a single objective and eyepiece lens. In contrast to the microscope, the objective lens of the telescope was designed with a large focal length to avoid optical aberrations. The objective focuses an image of a distant object at its focal point which is adjusted to be at the focal point of an eyepiece of a much smaller focal length. The main goal of a telescope is not necessarily magnification, but rather collection of light which is determined by the physical size of the objective lens. Thus, telescopes are normally indicated by the diameters of their objectives rather than by the magnification which can be changed by switching eyepieces. Because the magnification of a telescope is equal to the focal length of the objective divided by the focal length of the eyepiece, smaller focal-length eyepieces cause greater magnification.[83]
    Since crafting large lenses is much more difficult than crafting large mirrors, most modern telescopes are reflecting telescopes, that is, telescopes that use a primary mirror rather than an objective lens. The same general optical considerations apply to reflecting telescopes that applied to refracting telescopes, namely, the larger the primary mirror, the more light collected, and the magnification is still equal to the focal length of the primary mirror divided by the focal length of the eyepiece. Professional telescopes generally do not have eyepieces and instead place an instrument (often a charge-coupled device) at the focal point instead.[83]

    Photography


    Photograph taken with aperture f/32

    Photograph taken with aperture f/5
    The optics of photography involves both lenses and the medium in which the electromagnetic radiation is recorded, whether it be a plate, film, or charge-coupled device. Photographers must consider the reciprocity of the camera and the shot which is summarized by the relation
    Exposure ∝ ApertureArea × ExposureTime × SceneLuminance[85]
    In other words, the smaller the aperture (giving greater depth of focus), the less light coming in, so the length of time has to be increased (leading to possible blurriness if motion occurs). An example of the use of the law of reciprocity is the Sunny 16 rule which gives a rough estimate for the settings needed to estimate the proper exposure in daylight.[86]
    A camera's aperture is measured by a unitless number called the f-number or f-stop, f/#, often notated as N, and given by
    f/\# = N = \frac fD \
    where f is the focal length, and D is the diameter of the entrance pupil. By convention, "f/#" is treated as a single symbol, and specific values of f/# are written by replacing the number sign with the value. The two ways to increase the f-stop are to either decrease the diameter of the entrance pupil or change to a longer focal length (in the case of a zoom lens, this can be done by simply adjusting the lens). Higher f-numbers also have a larger depth of field due to the lens approaching the limit of a pinhole camera which is able to focus all images perfectly, regardless of distance, but requires very long exposure times.[87]
    The field of view that the lens will provide changes with the focal length of the lens. There are three basic classifications based on the relationship to the diagonal size of the film or sensor size of the camera to the focal length of the lens:[88]
    • Normal lens: angle of view of about 50° (called normal because this angle considered roughly equivalent to human vision[88]) and a focal length approximately equal to the diagonal of the film or sensor.[89]
    • Wide-angle lens: angle of view wider than 60° and focal length shorter than a normal lens.[90]
    • Long focus lens: angle of view narrower than a normal lens. This is any lens with a focal length longer than the diagonal measure of the film or sensor.[91] The most common type of long focus lens is the telephoto lens, a design that uses a special telephoto group to be physically shorter than its focal length.[92]
    Modern zoom lenses may have some or all of these attributes.
    The absolute value for the exposure time required depends on how sensitive to light the medium being used is (measured by the film speed, or, for digital media, by the quantum efficiency).[93] Early photography used media that had very low light sensitivity, and so exposure times had to be long even for very bright shots. As technology has improved, so has the sensitivity through film cameras and digital cameras.[94]
    Other results from physical and geometrical optics apply to camera optics. For example, the maximum resolution capability of a particular camera set-up is determined by the diffraction limit associated with the pupil size and given, roughly, by the Rayleigh criterion.[95]

    Atmospheric optics


    A colourful sky is often due to scattering of light off particulates and pollution, as in this photograph of a sunset during the October 2007 California wildfires.
    The unique optical properties of the atmosphere cause a wide range of spectacular optical phenomena. The blue colour of the sky is a direct result of Rayleigh scattering which redirects higher frequency (blue) sunlight back into the field of view of the observer. Because blue light is scattered more easily than red light, the sun takes on a reddish hue when it is observed through a thick atmosphere, as during a sunrise or sunset. Additional particulate matter in the sky can scatter different colours at different angles creating colourful glowing skies at dusk and dawn. Scattering off of ice crystals and other particles in the atmosphere are responsible for halos, afterglows, coronas, rays of sunlight, and sun dogs. The variation in these kinds of phenomena is due to different particle sizes and geometries.[96]
    Mirages are optical phenomena in which light rays are bent due to thermal variations in the refraction index of air, producing displaced or heavily distorted images of distant objects. Other dramatic optical phenomena associated with this include the Novaya Zemlya effect where the sun appears to rise earlier than predicted with a distorted shape. A spectacular form of refraction occurs with a temperature inversion called the Fata Morgana where objects on the horizon or even beyond the horizon, such as islands, cliffs, ships or icebergs, appear elongated and elevated, like "fairy tale castles".[97]
    Rainbows are the result of a combination of internal reflection and dispersive refraction of light in raindrops. A single reflection off the backs of an array of raindrops produces a rainbow with an angular size on the sky that ranges from 40° to 42° with red on the outside. Double rainbows are produced by two internal reflections with angular size of 50.5° to 54° with violet on the outside. Because rainbows are seen with the sun 180° away from the centre of the rainbow, rainbows are more prominent the closer the sun is to the horizon.[64]

    close

    Special relativity

    From Wikipedia, the free encyclopedia
      (Redirected from Special Relativity)
    In physics, special relativity (SR, also known as the special theory of relativity or STR) is the accepted physical theory regarding the relationship between space and time. It is based on two postulates: (1) that the laws of physics are invariant (i.e., identical) in all inertial systems (non-accelerating frames of reference); and (2) that the speed of light in a vacuum is the same for all observers, regardless of the motion of the light source. It was originally proposed in 1905 by Albert Einstein in the paper "On the Electrodynamics of Moving Bodies".[1] The inconsistency of classical mechanics with Maxwell’s equations of electromagnetism led to the development of special relativity, which corrects classical mechanics to handle situations involving motions nearing the speed of light. As of today, special relativity is the most accurate model of motion at any speed. Even so, classical mechanics is still useful (due to its simplicity and high accuracy) as an approximation at small velocities relative to the speed of light.
    Special relativity implies a wide range of consequences, which have been experimentally verified,[2] including length contraction, time dilation, relativistic mass, mass–energy equivalence, a universal speed limit, and relativity of simultaneity. It has replaced the conventional notion of an absolute universal time with the notion of a time that is dependent on reference frame and spatial position. Rather than an invariant time interval between two events, there is an invariant spacetime interval. Combined with other laws of physics, the two postulates of special relativity predict the equivalence of mass and energy, as expressed in the mass–energy equivalence formula E = mc2, where c is the speed of light in vacuum.[3][4]
    A defining feature of special relativity is the replacement of the Galilean transformations of classical mechanics with the Lorentz transformations. Time and space cannot be defined separately from one another. Rather space and time are interwoven into a single continuum known as spacetime. Events that occur at the same time for one observer could occur at different times for another.
    The theory is called "special" because it applied the principle of relativity only to the special case of inertial reference frames. Einstein later published a paper on general relativity in 1915 to apply the principle in the general case, that is, to any frame so as to handle general coordinate transformations, and gravitational effects.
    As Galilean relativity is now considered an approximation of special relativity valid for low speeds, special relativity is considered an approximation of the theory of general relativity valid for weak gravitational fields. The presence of gravity becomes undetectable at sufficiently small-scale, free-falling conditions. General relativity incorporates noneuclidean geometry, so that the gravitational effects are represented by the geometric curvature of spacetime. Contrarily, special relativity is restricted to flat spacetime. The geometry of spacetime in special relativity is called Minkowski space. A locally Lorentz invariant frame that abides by Special relativity can be defined at sufficiently small scales, even in curved spacetime.
    Galileo Galilei had already postulated that there is no absolute and well-defined state of rest (no privileged reference frames), a principle now called Galileo's principle of relativity. Einstein extended this principle so that it accounted for the constant speed of light,[5] a phenomenon that had been recently observed in the Michelson–Morley experiment. He also postulated that it holds for all the laws of physics, including both the laws of mechanics and of electrodynamics.[6]

    Albert Einstein around 1905

    Postulates

    Reflections of this type made it clear to me as long ago as shortly after 1900, i.e., shortly after Planck's trailblazing work, that neither mechanics nor electrodynamics could (except in limiting cases) claim exact validity. Gradually I despaired of the possibility of discovering the true laws by means of constructive efforts based on known facts. The longer and the more desperately I tried, the more I came to the conviction that only the discovery of a universal formal principle could lead us to assured results... How, then, could such a universal principle be found?
    —Albert Einstein: Autobiographical Notes[7]
    Einstein discerned two fundamental propositions that seemed to be the most assured, regardless of the exact validity of the (then) known laws of either mechanics or electrodynamics. These propositions were the constancy of the speed of light and the independence of physical laws (especially the constancy of the speed of light) from the choice of inertial system. In his initial presentation of special relativity in 1905 he expressed these postulates as:[1]
    • The Principle of Relativity – The laws by which the states of physical systems undergo change are not affected, whether these changes of state be referred to the one or the other of two systems in uniform translatory motion relative to each other.[1]
    • The Principle of Invariant Light Speed – "... light is always propagated in empty space with a definite velocity [speed] c which is independent of the state of motion of the emitting body." (from the preface).[1] That is, light in vacuum propagates with the speed c (a fixed constant, independent of direction) in at least one system of inertial coordinates (the "stationary system"), regardless of the state of motion of the light source.
    The derivation of special relativity depends not only on these two explicit postulates, but also on several tacit assumptions (made in almost all theories of physics), including the isotropy and homogeneity of space and the independence of measuring rods and clocks from their past history.[8]
    Following Einstein's original presentation of special relativity in 1905, many different sets of postulates have been proposed in various alternative derivations.[9] However, the most common set of postulates remains those employed by Einstein in his original paper. A more mathematical statement of the Principle of Relativity made later by Einstein, which introduces the concept of simplicity not mentioned above is:
    Special principle of relativity: If a system of coordinates K is chosen so that, in relation to it, physical laws hold good in their simplest form, the same laws hold good in relation to any other system of coordinates K' moving in uniform translation relatively to K.[10]
    Henri Poincaré provided the mathematical framework for relativity theory by proving that Lorentz transformations are a subset of his Poincaré group of symmetry transformations. Einstein later derived these transformations from his axioms.
    Many of Einstein's papers present derivations of the Lorentz transformation based upon these two principles.[11]
    Einstein consistently based the derivation of Lorentz invariance (the essential core of special relativity) on just the two basic principles of relativity and light-speed invariance. He wrote:
    The insight fundamental for the special theory of relativity is this: The assumptions relativity and light speed invariance are compatible if relations of a new type ("Lorentz transformation") are postulated for the conversion of coordinates and times of events... The universal principle of the special theory of relativity is contained in the postulate: The laws of physics are invariant with respect to Lorentz transformations (for the transition from one inertial system to any other arbitrarily chosen inertial system). This is a restricting principle for natural laws...[7]
    Thus many modern treatments of special relativity base it on the single postulate of universal Lorentz covariance, or, equivalently, on the single postulate of Minkowski spacetime.[12][13]
    From the principle of relativity alone without assuming the constancy of the speed of light (i.e. using the isotropy of space and the symmetry implied by the principle of special relativity) one can show that the spacetime transformations between inertial frames are either Euclidean, Galilean, or Lorentzian. In the Lorentzian case, one can then obtain relativistic interval conservation and a certain finite limiting speed. Experiments suggest that this speed is the speed of light in vacuum.[14][15]
    The constancy of the speed of light was motivated by Maxwell's theory of electromagnetism and the lack of evidence for the luminiferous ether. There is conflicting evidence on the extent to which Einstein was influenced by the null result of the Michelson–Morley experiment.[16][17] In any case, the null result of the Michelson–Morley experiment helped the notion of the constancy of the speed of light gain widespread and rapid acceptance.

    Lack of an absolute reference frame

    The principle of relativity, which states that there is no preferred inertial reference frame, dates back to Galileo, and was incorporated into Newtonian physics. However, in the late 19th century, the existence of electromagnetic waves led physicists to suggest that the universe was filled with a substance that they called "aether", which would act as the medium through which these waves, or vibrations travelled. The aether was thought to constitute an absolute reference frame against which speeds could be measured, and could be considered fixed and motionless. Aether supposedly possessed some wonderful properties: it was sufficiently elastic to support electromagnetic waves, and those waves could interact with matter, yet it offered no resistance to bodies passing through it. The results of various experiments, including the Michelson–Morley experiment, indicated that the Earth was always 'stationary' relative to the aether – something that was difficult to explain, since the Earth is in orbit around the Sun. Einstein's solution was to discard the notion of an aether and the absolute state of rest. In relativity, any reference frame moving with uniform motion will observe the same laws of physics. In particular, the speed of light in vacuum is always measured to be c, even when measured by multiple systems that are moving at different (but constant) velocities.

    Reference frames, coordinates and the Lorentz transformation


    The primed system is in motion relative to the unprimed system with constant speed v only along the x-axis, from the perspective of an observer stationary in the unprimed system. By the principle of relativity, an observer stationary in the primed system will view a likewise construction except that the speed they record will be −v. The changing of the speed of propagation of interaction from infinite in non-relativistic mechanics to a finite value will require a modification of the transformation equations mapping events in one frame to another.
    Relativity theory depends on "reference frames". The term reference frame as used here is an observational perspective in space which is not undergoing any change in motion (acceleration), from which a position can be measured along 3 spatial axes. In addition, a reference frame has the ability to determine measurements of the time of events using a 'clock' (any reference device with uniform periodicity).
    An event is an occurrence that can be assigned a single unique time and location in space relative to a reference frame: it is a "point" in spacetime. Since the speed of light is constant in relativity in each and every reference frame, pulses of light can be used to unambiguously measure distances and refer back the times that events occurred to the clock, even though light takes time to reach the clock after the event has transpired.
    For example, the explosion of a firecracker may be considered to be an "event". We can completely specify an event by its four spacetime coordinates: The time of occurrence and its 3-dimensional spatial location define a reference point. Let's call this reference frame S.
    In relativity theory we often want to calculate the position of a point from a different reference point.
    Suppose we have a second reference frame S′, whose spatial axes and clock exactly coincide with that of S at time zero, but it is moving at a constant velocity v with respect to S along the x-axis.
    Since there is no absolute reference frame in relativity theory, a concept of 'moving' doesn't strictly exist, as everything is always moving with respect to some other reference frame. Instead, any two frames that move at the same speed in the same direction are said to be comoving. Therefore S and S′ are not comoving.
    Define the event to have spacetime coordinates (t,x,y,z) in system S and (t′,x′,y′,z′) in S′. Then the Lorentz transformation specifies that these coordinates are related in the following way:
    \begin{align}
t' &= \gamma \ (t - vx/c^2) \\
x' &= \gamma \ (x - v t) \\
y' &= y \\
z' &= z ,
\end{align}
    where
    \gamma = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}}
    is the Lorentz factor and c is the speed of light in vacuum, and the velocity v of S′ is parallel to the x-axis. The y and z coordinates are unaffected; only the x and t coordinates are transformed. These Lorentz transformations form a one-parameter group of linear mappings, that parameter being called rapidity.
    There is nothing special about the x-axis, the transformation can apply to the y or z axes, or indeed in any direction, which can be done by directions parallel to the motion (which are warped by the γ factor) and perpendicular; see main article for details.
    A quantity invariant under Lorentz transformations is known as a Lorentz scalar.
    Writing the Lorentz transformation and its inverse in terms of coordinate differences, where for instance one event has coordinates (x1, t1) and (x1, t1), another event has coordinates (x2, t2) and (x2, t2), and the differences are defined as
     \begin{array}{ll}
\Delta x' = x'_2-x'_1 \ , & \Delta x = x_2-x_1 \ , \\
\Delta t' = t'_2-t'_1 \ , & \Delta t = t_2-t_1 \ , \\
\end{array}
    we get
     \begin{array}{ll}
\Delta x' = \gamma \ (\Delta x - v \,\Delta t) \ , & \Delta x = \gamma \ (\Delta x' + v \,\Delta t') \ , \\
\Delta t' = \gamma \ \left(\Delta t - \dfrac{v \,\Delta x}{c^{2}} \right) \ , & \Delta t = \gamma \ \left(\Delta t' + \dfrac{v \,\Delta x'}{c^{2}} \right) \ . \\
\end{array}
    These effects are not merely appearances; they are explicitly related to our way of measuring time intervals between events which occur at the same place in a given coordinate system (called "co-local" events). These time intervals will be different in another coordinate system moving with respect to the first, unless the events are also simultaneous. Similarly, these effects also relate to our measured distances between separated but simultaneous events in a given coordinate system of choice. If these events are not co-local, but are separated by distance (space), they will not occur at the same spatial distance from each other when seen from another moving coordinate system. However, the spacetime interval will be the same for all observers. The underlying reality remains the same. Only our perspective changes.

    Consequences derived from the Lorentz transformation

    The consequences of special relativity can be derived from the Lorentz transformation equations.[18] These transformations, and hence special relativity, lead to different physical predictions than those of Newtonian mechanics when relative velocities become comparable to the speed of light. The speed of light is so much larger than anything humans encounter that some of the effects predicted by relativity are initially counterintuitive.

    Relativity of simultaneity


    Event B is simultaneous with A in the green reference frame, but it occurred before in the blue frame, and will occur later in the red frame.
    Two events happening in two different locations that occur simultaneously in the reference frame of one inertial observer, may occur non-simultaneously in the reference frame of another inertial observer (lack of absolute simultaneity).
    From the first equation of the Lorentz transformation in terms of coordinate differences
    \Delta t' = \gamma \left(\Delta t - \frac{v \,\Delta x}{c^{2}} \right)
    it is clear that two events that are simultaneous in frame S (satisfying Δt = 0), are not necessarily simultaneous in another inertial frame S′ (satisfying Δt′ = 0). Only if these events are additionally co-local in frame S (satisfying Δx = 0), will they be simultaneous in another frame S′.

    Time dilation

    The time lapse between two events is not invariant from one observer to another, but is dependent on the relative speeds of the observers' reference frames (e.g., the twin paradox which concerns a twin who flies off in a spaceship traveling near the speed of light and returns to discover that his or her twin sibling has aged much more).
    Suppose a clock is at rest in the unprimed system S. Two different ticks of this clock are then characterized by Δx = 0. To find the relation between the times between these ticks as measured in both systems, the first equation can be used to find:
    \Delta t' = \gamma\, \Delta t     for events satisfying    \Delta x = 0 \ .
    This shows that the time (Δt') between the two ticks as seen in the frame in which the clock is moving (S′), is longer than the time (Δt) between these ticks as measured in the rest frame of the clock (S). Time dilation explains a number of physical phenomena; for example, the decay rate of muons produced by cosmic rays impinging on the Earth's atmosphere.[19]

    Length contraction

    The dimensions (e.g., length) of an object as measured by one observer may be smaller than the results of measurements of the same object made by another observer (e.g., the ladder paradox involves a long ladder traveling near the speed of light and being contained within a smaller garage).
    Similarly, suppose a measuring rod is at rest and aligned along the x-axis in the unprimed system S. In this system, the length of this rod is written as Δx. To measure the length of this rod in the system S′, in which the clock is moving, the distances x′ to the end points of the rod must be measured simultaneously in that system S′. In other words, the measurement is characterized by Δt′ = 0, which can be combined with the fourth equation to find the relation between the lengths Δx and Δx′:
    \Delta x' = \frac{\Delta x}{\gamma}     for events satisfying    \Delta t' = 0 \ .
    This shows that the length (Δx′) of the rod as measured in the frame in which it is moving (S′), is shorter than its length (Δx) in its own rest frame (S).

    Composition of velocities

    Velocities (speeds) do not simply add. If the observer in S measures an object moving along the x axis at velocity u, then the observer in the S′ system, a frame of reference moving at velocity v in the x direction with respect to S, will measure the object moving with velocity u′ where (from the Lorentz transformations above):
    u'=\frac{dx'}{dt'}=\frac{\gamma \ (dx-v dt)}{\gamma \ (dt-v dx/c^2)}=\frac{(dx/dt)-v}{1-(v/c^2)(dx/dt)}=\frac{u-v}{1-uv/c^2} \ .
    The other frame S will measure:
    u=\frac{dx}{dt}=\frac{\gamma \ (dx'+v dt')}{\gamma \ (dt'+v dx'/c^2)}=\frac{(dx'/dt')+v}{1+(v/c^2)(dx'/dt')}=\frac{u'+v}{1+u'v/c^2} \ .
    Notice that if the object were moving at the speed of light in the S system (i.e. u = c), then it would also be moving at the speed of light in the S′ system. Also, if both u and v are small with respect to the speed of light, we will recover the intuitive Galilean transformation of velocities
    u' \approx u-v \ .
    The usual example given is that of a train (frame S′ above) traveling due east with a velocity v with respect to the tracks (frame S). A child inside the train throws a baseball due east with a velocity u′ with respect to the train. In classical physics, an observer at rest on the tracks will measure the velocity of the baseball (due east) as u = u′ + v, while in special relativity this is no longer true; instead the velocity of the baseball (due east) is given by the second equation: u = (u′ + v)/(1 + uv/c2). Again, there is nothing special about the x or east directions. This formalism applies to any direction by considering parallel and perpendicular motion to the direction of relative velocity v, see main article for details.
    Einstein's addition of colinear velocities is consistent with the Fizeau experiment which determined the speed of light in a fluid moving parallel to the light, but no experiment has ever tested the formula for the general case of non-parallel velocities.[citation needed]

    Other consequences

    Thomas rotation

    The orientation of an object (i.e. the alignment of its axes with the observer's axes) may be different for different observers. Unlike other relativistic effects, this effect becomes quite significant at fairly low velocities as can be seen in the spin of moving particles.

    Equivalence of mass and energy

    As an object's speed approaches the speed of light from an observer's point of view, its relativistic mass increases thereby making it more and more difficult to accelerate it from within the observer's frame of reference.
    The energy content of an object at rest with mass m equals mc2. Conservation of energy implies that, in any reaction, a decrease of the sum of the masses of particles must be accompanied by an increase in kinetic energies of the particles after the reaction. Similarly, the mass of an object can be increased by taking in kinetic energies.
    In addition to the papers referenced above—which give derivations of the Lorentz transformation and describe the foundations of special relativity—Einstein also wrote at least four papers giving heuristic arguments for the equivalence (and transmutability) of mass and energy, for E = mc2.
    Mass–energy equivalence is a consequence of special relativity. The energy and momentum, which are separate in Newtonian mechanics, form a four-vector in relativity, and this relates the time component (the energy) to the space components (the momentum) in a nontrivial way. For an object at rest, the energy–momentum four-vector is (E, 0, 0, 0): it has a time component which is the energy, and three space components which are zero. By changing frames with a Lorentz transformation in the x direction with a small value of the velocity v, the energy momentum four-vector becomes (E, Ev/c2, 0, 0). The momentum is equal to the energy multiplied by the velocity divided by c2. As such, the Newtonian mass of an object, which is the ratio of the momentum to the velocity for slow velocities, is equal to E/c2.
    The energy and momentum are properties of matter and radiation, and it is impossible to deduce that they form a four-vector just from the two basic postulates of special relativity by themselves, because these don't talk about matter or radiation, they only talk about space and time. The derivation therefore requires some additional physical reasoning. In his 1905 paper, Einstein used the additional principles that Newtonian mechanics should hold for slow velocities, so that there is one energy scalar and one three-vector momentum at slow velocities, and that the conservation law for energy and momentum is exactly true in relativity. Furthermore, he assumed that the energy of light is transformed by the same Doppler-shift factor as its frequency, which he had previously shown to be true based on Maxwell's equations.[1] The first of Einstein's papers on this subject was "Does the Inertia of a Body Depend upon its Energy Content?" in 1905.[20] Although Einstein's argument in this paper is nearly universally accepted by physicists as correct, even self-evident, many authors over the years have suggested that it is wrong.[21] Other authors suggest that the argument was merely inconclusive because it relied on some implicit assumptions.[22]
    Einstein acknowledged the controversy over his derivation in his 1907 survey paper on special relativity. There he notes that it is problematic to rely on Maxwell's equations for the heuristic mass–energy argument. The argument in his 1905 paper can be carried out with the emission of any massless particles, but the Maxwell equations are implicitly used to make it obvious that the emission of light in particular can be achieved only by doing work. To emit electromagnetic waves, all you have to do is shake a charged particle, and this is clearly doing work, so that the emission is of energy.[23][24]

    How far can one travel from the Earth?

    Since one can not travel faster than light, one might conclude that a human can never travel further from Earth than 40 light years if the traveler is active between the age of 20 and 60. One would easily think that a traveler would never be able to reach more than the very few solar systems which exist within the limit of 20–40 light years from the earth. But that would be a mistaken conclusion. Because of time dilation, a hypothetical spaceship can travel thousands of light years during the pilot's 40 active years. If a spaceship could be built that accelerates at a constant 1g, it will after a little less than a year be traveling at almost the speed of light as seen from Earth. Time dilation will increase his life span as seen from the reference system of the Earth, but his lifespan measured by a clock traveling with him will not thereby change. During his journey, people on Earth will experience more time than he does. A 5 year round trip for him will take 6½ Earth years and cover a distance of over 6 light-years. A 20 year round trip for him (5 years accelerating, 5 decelerating, twice each) will land him back on Earth having traveled for 335 Earth years and a distance of 331 light years.[25] A full 40 year trip at 1 g will appear on Earth to last 58,000 years and cover a distance of 55,000 light years. A 40 year trip at 1.1 g will take 148,000 Earth years and cover about 140,000 light years. A one-way 28 year (14 years accelerating, 14 decelerating as measured with the cosmonaut's clock) trip at 1 g acceleration could reach 2,000,000 light-years to the Andromeda Galaxy.[26] This same time dilation is why a muon traveling close to c is observed to travel much further than c times its half-life (when at rest).[27]

    Causality and prohibition of motion faster than light


    Diagram 2. Light cone
    In diagram 2 the interval AB is 'time-like'; i.e., there is a frame of reference in which events A and B occur at the same location in space, separated only by occurring at different times. If A precedes B in that frame, then A precedes B in all frames. It is hypothetically possible for matter (or information) to travel from A to B, so there can be a causal relationship (with A the cause and B the effect).
    The interval AC in the diagram is 'space-like'; i.e., there is a frame of reference in which events A and C occur simultaneously, separated only in space. There are also frames in which A precedes C (as shown) and frames in which C precedes A. If it were possible for a cause-and-effect relationship to exist between events A and C, then paradoxes of causality would result. For example, if A was the cause, and C the effect, then there would be frames of reference in which the effect preceded the cause. Although this in itself won't give rise to a paradox, one can show[28][29] that faster than light signals can be sent back into one's own past. A causal paradox can then be constructed by sending the signal if and only if no signal was received previously.
    Therefore, if causality is to be preserved, one of the consequences of special relativity is that no information signal or material object can travel faster than light in vacuum. However, some "things" can still move faster than light. For example, the location where the beam of a search light hits the bottom of a cloud can move faster than light when the search light is turned rapidly.[30]
    Even without considerations of causality, there are other strong reasons why faster-than-light travel is forbidden by special relativity. For example, if a constant force is applied to an object for a limitless amount of time, then integrating F = dp/dt gives a momentum that grows without bound, but this is simply because p = m \gamma v approaches infinity as v approaches c. To an observer who is not accelerating, it appears as though the object's inertia is increasing, so as to produce a smaller acceleration in response to the same force. This behavior is observed in particle accelerators, where each charged particle is accelerated by the electromagnetic force.
    Theoretical and experimental tunneling studies carried out by Günter Nimtz and Petrissa Eckle claimed that under special conditions signals may travel faster than light.[31][32][33][34] It was measured that fiber digital signals were traveling up to 5 times c and a zero-time tunneling electron carried the information that the atom is ionized, with photons, phonons and electrons spending zero time in the tunneling barrier. According to Nimtz and Eckle, in this superluminal process only the Einstein causality and the special relativity but not the primitive causality are violated: Superluminal propagation does not result in any kind of time travel.[35][36] Several scientists have stated not only that Nimtz' interpretations were erroneous, but also that the experiment actually provided a trivial experimental confirmation of the special relativity theory.[37][38][39]

    Geometry of spacetime

    Comparison between flat Euclidean space and Minkowski space


    Orthogonality and rotation of coordinate systems compared between left: Euclidean space through circular angle φ, right: in Minkowski spacetime through hyperbolic angle φ (red lines labelled c denote the worldlines of a light signal, a vector is orthogonal to itself if it lies on this line).[40]
    Special relativity uses a 'flat' 4-dimensional Minkowski space – an example of a spacetime. Minkowski spacetime appears to be very similar to the standard 3-dimensional Euclidean space, but there is a crucial difference with respect to time.
    In 3D space, the differential of distance (line element) ds is defined by
     ds^2 = d\mathbf{x} \cdot d\mathbf{x} = dx_1^2 + dx_2^2 + dx_3^2,
    where dx = (dx1, dx2, dx3) are the differentials of the three spatial dimensions. In Minkowski geometry, there is an extra dimension with coordinate X0 derived from time, such that the distance differential fulfills
     ds^2 = -dX_0^2 + dX_1^2 + dX_2^2 + dX_3^2,
    where dX = (dX0, dX1, dX2, dX3) are the differentials of the four spacetime dimensions. This suggests a deep theoretical insight: special relativity is simply a rotational symmetry of our spacetime, analogous to the rotational symmetry of Euclidean space (see image right).[41] Just as Euclidean space uses a Euclidean metric, so spacetime uses a Minkowski metric. Basically, special relativity can be stated as the invariance of any spacetime interval (that is the 4D distance between any two events) when viewed from any inertial reference frame. All equations and effects of special relativity can be derived from this rotational symmetry (the Poincaré group) of Minkowski spacetime.
    The actual form of ds above depends on the metric and on the choices for the X0 coordinate. To make the time coordinate look like the space coordinates, it can be treated as imaginary: X0 = ict (this is called a Wick rotation). According to Misner, Thorne and Wheeler (1971, §2.3), ultimately the deeper understanding of both special and general relativity will come from the study of the Minkowski metric (described below) and to take X0 = ct, rather than a "disguised" Euclidean metric using ict as the time coordinate.
    Some authors use X0 = t, with factors of c elsewhere to compensate; for instance, spatial coordinates are divided by c or factors of c±2 are included in the metric tensor.[42] These numerous conventions can be superseded by using natural units where c = 1. Then space and time have equivalent units, and no factors of c appear anywhere.

    3D spacetime


    Three dimensional dual-cone.

    Null spherical space.
    If we reduce the spatial dimensions to 2, so that we can represent the physics in a 3D space
     ds^2 = dx_1^2 + dx_2^2 - c^2 dt^2,
    we see that the null geodesics lie along a dual-cone (see image right) defined by the equation;
     ds^2 = 0 = dx_1^2 + dx_2^2 - c^2 dt^2
    or simply
     dx_1^2 + dx_2^2 = c^2 dt^2,
     which is the equation of a circle of radius c dt.

    4D spacetime

    If we extend this to three spatial dimensions, the null geodesics are the 4-dimensional cone:
     ds^2 = 0 = dx_1^2 + dx_2^2 + dx_3^2 - c^2 dt^2
    so
     dx_1^2 + dx_2^2 + dx_3^2 = c^2 dt^2.
    This null dual-cone represents the "line of sight" of a point in space. That is, when we look at the stars and say "The light from that star which I am receiving is X years old", we are looking down this line of sight: a null geodesic. We are looking at an event a distance d = \sqrt{x_1^2+x_2^2+x_3^2} away and a time d/c in the past. For this reason the null dual cone is also known as the 'light cone'. (The point in the lower left of the picture below represents the star, the origin represents the observer, and the line represents the null geodesic "line of sight".)
    The cone in the −t region is the information that the point is 'receiving', while the cone in the +t section is the information that the point is 'sending'.
    The geometry of Minkowski space can be depicted using Minkowski diagrams, which are useful also in understanding many of the thought-experiments in special relativity.
    Note that, in 4d spacetime, the concept of the center of mass becomes more complicated, see center of mass (relativistic).

    Physics in spacetime

    Transformations of physical quantities between reference frames

    Above, the Lorentz transformation for the time coordinate and three space coordinates illustrates that they are intertwined. This is true more generally: certain pairs of "timelike" and "spacelike" quantities naturally combine on equal footing under the same Lorentz transformation.
    The Lorentz transformation in standard configuration above, i.e. for a boost in the x direction, can be recast into matrix form as follows:
    \begin{pmatrix}
ct'\\ x'\\ y'\\ z'
\end{pmatrix} = \begin{pmatrix}
\gamma & -\beta\gamma & 0 & 0\\
-\beta\gamma & \gamma & 0 & 0\\
0 & 0 & 1 & 0\\
0 & 0 & 0 & 1
\end{pmatrix}
\begin{pmatrix}
ct\\ x\\ y\\ z
\end{pmatrix} =
\begin{pmatrix}
\gamma ct- \gamma\beta x\\
\gamma x - \beta \gamma ct \\ y\\ z
\end{pmatrix}.
    In Newtonian mechanics, quantities which have magnitude and direction are mathematically described as 3d vectors in Euclidean space, and in general they are parametrized by time. In special relativity, this notion is extended by adding the appropriate timelike quantity to a spacelike vector quantity, and we have 4d vectors, or "four vectors", in Minkowski spacetime. The components of vectors are written using tensor index notation, as this has numerous advantages. The notation makes it clear the equations are manifestly covariant under the Poincaré group, thus bypassing the tedious calculations to check this fact. In constructing such equations, we often find that equations previously thought to be unrelated are, in fact, closely connected being part of the same tensor equation. Recognizing other physical quantities as tensors simplifies their transformation laws. Throughout, upper indices (superscripts) are contravariant indices rather than exponents except when they indicate a square (this is should be clear from the context), and lower indices (subscripts) are covariant indices. For simplicity and consistency with the earlier equations, Cartesian coordinates will be used.
    The simplest example of a four-vector is the position of an event in spacetime, which constitutes a timelike component ct and spacelike component x = (x, y, z), in a contravariant position four vector with components:
    X^\nu = (X^0, X^1, X^2, X^3)= (ct, x, y, z).
    where we define X0 = ct so that the time coordinate has the same dimension of distance as the other spatial dimensions; so that space and time are treated equally.[43][44][45] Now the transformation of the contravariant components of the position 4-vector can be compactly written as:
    X^{\mu'}=\Lambda^{\mu'}{}_\nu X^\nu
    where there is an implied summation on ν from 0 to 3, and \Lambda^{\mu'}{}_{\nu} is a matrix.
    More generally, all contravariant components of a four-vector T^\nu transform from one frame to another frame by a Lorentz transformation:
    T^{\mu'} = \Lambda^{\mu'}{}_{\nu} T^\nu
    Examples of other 4-vectors include the four-velocity Uμ, defined as the derivative of the position 4-vector with respect to proper time:
    U^\mu = \frac{dX^\mu}{d\tau} = \gamma(v)( c , v_x , v_y, v_z ) .
    where the Lorentz factor is:
    \gamma(v)= \frac{1}{\sqrt{1- (v/c)^2}} \,,\quad v^2 = v_x^2 + v_y^2 + v_z^2 \,.
    The relativistic energy E = \gamma(v)mc^2 and relativistic momentum \mathbf{p} = \gamma(v)m \mathbf{v} of an object are respectively the timelike and spacelike components of a covariant four momentum vector:
    P_\nu = m U_\nu = m\gamma(v)(c,v_x,v_y,v_z)= (E/c,p_x,p_y,p_z).
    where m is the invariant mass.
    The four-acceleration is the proper time derivative of 4-velocity:
    A^\mu = \frac{d U^\mu}{d\tau} \,.
    The transformation rules for three-dimensional velocities and accelerations are very awkward; even above in standard configuration the velocity equations are quite complicated owing to their non-linearity. On the other hand, the transformation of four-velocity and four-acceleration are simpler by means of the Lorentz transformation matrix.
    The four-gradient of a scalar field φ transforms covariantly rather than contravariantly:
    \begin{pmatrix} \frac{1}{c}\frac{\partial \phi}{\partial t'} & \frac{\partial \phi}{\partial x'} & \frac{\partial \phi}{\partial y'} & \frac{\partial \phi}{\partial z'}\end{pmatrix} = \begin{pmatrix} \frac{1}{c}\frac{\partial \phi}{\partial t} & \frac{\partial \phi}{\partial x} & \frac{\partial \phi}{\partial y} & \frac{\partial \phi}{\partial z}\end{pmatrix}\begin{pmatrix}
\gamma & -\beta\gamma & 0 & 0\\
-\beta\gamma & \gamma & 0 & 0\\
0 & 0 & 1 & 0\\
0 & 0 & 0 & 1
\end{pmatrix} \,.
    that is:
    (\partial_{\mu'} \phi) = \Lambda_{\mu'}{}^{\nu} (\partial_\nu \phi)\,,\quad \partial_{\mu} \equiv \frac{\partial}{\partial x^{\mu}}\,.
    only in Cartesian coordinates. It's the covariant derivative which transforms in manifest covariance, in Cartesian coordinates this happens to reduce to the partial derivatives, but not in other coordinates.
    More generally, the covariant components of a 4-vector transform according to the inverse Lorentz transformation:
    \Lambda_{\mu'}{}^{\nu} T^{\mu'} =  T^\nu
    where  \Lambda_{\mu'}{}^{\nu} is the reciprocal matrix of \Lambda^{\mu'}{}_{\nu}.
    The postulates of special relativity constrain the exact form the Lorentz transformation matrices take.
    More generally, most physical quantities are best described as (components of) tensors. So to transform from one frame to another, we use the well-known tensor transformation law[46]
    T^{\alpha' \beta' \cdots \zeta'}_{\theta' \iota' \cdots \kappa'} =
\Lambda^{\alpha'}{}_{\mu} \Lambda^{\beta'}{}_{\nu} \cdots \Lambda^{\zeta'}{}_{\rho}
\Lambda_{\theta'}{}^{\sigma} \Lambda_{\iota'}{}^{\upsilon} \cdots \Lambda_{\kappa'}{}^{\phi}
T^{\mu \nu \cdots \rho}_{\sigma \upsilon \cdots \phi}
    where \Lambda_{\chi'}{}^{\psi} is the reciprocal matrix of \Lambda^{\chi'}{}_{\psi}. All tensors transform by this rule.
    An example of a four dimensional second order antisymmetric tensor is the relativistic angular momentum, which has six components: three are the classical angular momentum, and the other three are related to the boost of the center of mass of the system. The derivative of the relativistic angular momentum with respect to proper time is the relativistic torque, also second order antisymmetric tensor.
    The electromagnetic field tensor is another second order antisymmetric tensor field, with six components: three for the electric field and another three for the magnetic field. There is also the stress–energy tensor for the electromagnetic field, namely the electromagnetic stress–energy tensor.

    Metric

    The metric tensor allows one to define the inner product of two vectors, which in turn allows one to assign a magnitude to the vector. Given the four-dimensional nature of spacetime the Minkowski metric η has components (valid in any inertial reference frame) which can be arranged in a 4 × 4 matrix:
    \eta_{\alpha\beta} = \begin{pmatrix}
-1 & 0 & 0 & 0\\
0 & 1 & 0 & 0\\
0 & 0 & 1 & 0\\
0 & 0 & 0 & 1
\end{pmatrix}
    which is equal to its reciprocal, \eta^{\alpha\beta}, in those frames. Throughout we use the signs as above, different authors use different conventions – see Minkowski metric alternative signs.
    The Poincaré group is the most general group of transformations which preserves the Minkowski metric:
    \eta_{\alpha\beta} = \eta_{\mu'\nu'} \Lambda^{\mu'}{}_\alpha \Lambda^{\nu'}{}_\beta \!
    and this is the physical symmetry underlying special relativity.
    The metric can be used for raising and lowering indices on vectors and tensors. Invariants can be constructed using the metric, the inner product of a 4-vector T with another 4-vector S is:
    T^{\alpha}S_{\alpha}=T^{\alpha}\eta_{\alpha\beta}S^{\beta} = T_{\alpha}\eta^{\alpha\beta}S_{\beta} = \text{invariant scalar}
    Invariant means that it takes the same value in all inertial frames, because it is a scalar (0 rank tensor), and so no Λ appears in its trivial transformation. The magnitude of the 4-vector T is the positive square root of the inner product with itself:
    |\mathbf{T}| = \sqrt{T^{\alpha}T_{\alpha}}
    One can extend this idea to tensors of higher order, for a second order tensor we can form the invariants:
    T^{\alpha}{}_{\alpha}\,,T^{\alpha}{}_{\beta}T^{\beta}{}_{\alpha}\,,T^{\alpha}{}_{\beta}T^{\beta}{}_{\gamma}T^{\gamma}{}_{\alpha} = \text{invariant scalars}\,,
    similarly for higher order tensors. Invariant expressions, particularly inner products of 4-vectors with themselves, provide equations that are useful for calculations, because one doesn't need to perform Lorentz transformations to determine the invariants.

    Relativistic kinematics and invariance

    The coordinate differentials transform also contravariantly:
    dX^{\mu'}=\Lambda^{\mu'}{}_\nu dX^\nu
    so the squared length of the differential of the position four-vector dXμ constructed using
    d\mathbf{X}^2 = dX^\mu \,dX_\mu = \eta_{\mu\nu}\,dX^\mu \,dX^\nu = -(c dt)^2+(dx)^2+(dy)^2+(dz)^2\,
    is an invariant. Notice that when the line element dX2 is negative that dX2 is the differential of proper time, while when dX2 is positive, dX2 is differential of the proper distance.
    The 4-velocity Uμ has an invariant form:
    {\mathbf U}^2 = \eta_{\nu\mu} U^\nu U^\mu = -c^2 \,,
    which means all velocity four-vectors have a magnitude of c. This is an expression of the fact that there is no such thing as being at coordinate rest in relativity: at the least, you are always moving forward through time. Differentiating the above equation by τ produces:
    2\eta_{\mu\nu}A^\mu U^\nu = 0.
    So in special relativity, the acceleration four-vector and the velocity four-vector are orthogonal.

    Relativistic dynamics and invariance

    The invariant magnitude of the momentum 4-vector generates the energy–momentum relation:
    \mathbf{P}^2 = \eta^{\mu\nu}P_\mu P_\nu = -(E/c)^2 + p^2 .
    We can work out what this invariant is by first arguing that, since it is a scalar, it doesn't matter which reference frame we calculate it, and then by transforming to a frame where the total momentum is zero.
    \mathbf{P}^2 = - (E_\mathrm{rest}/c)^2 = - (m c)^2 .
    We see that the rest energy is an independent invariant. A rest energy can be calculated even for particles and systems in motion, by translating to a frame in which momentum is zero.
    The rest energy is related to the mass according to the celebrated equation discussed above:
    E_\mathrm{rest} = m c^2.
    Note that the mass of systems measured in their center of momentum frame (where total momentum is zero) is given by the total energy of the system in this frame. It may not be equal to the sum of individual system masses measured in other frames.
    To use Newton's third law of motion, both forces must be defined as the rate of change of momentum with respect to the same time coordinate. That is, it requires the 3D force defined above. Unfortunately, there is no tensor in 4D which contains the components of the 3D force vector among its components.
    If a particle is not traveling at c, one can transform the 3D force from the particle's co-moving reference frame into the observer's reference frame. This yields a 4-vector called the four-force. It is the rate of change of the above energy momentum four-vector with respect to proper time. The covariant version of the four-force is:
    F_\nu = \frac{d P_{\nu}}{d \tau} = m A_\nu
    In the rest frame of the object, the time component of the four force is zero unless the "invariant mass" of the object is changing (this requires a non-closed system in which energy/mass is being directly added or removed from the object) in which case it is the negative of that rate of change of mass, times c. In general, though, the components of the four force are not equal to the components of the three-force, because the three force is defined by the rate of change of momentum with respect to coordinate time, i.e. dp/dt while the four force is defined by the rate of change of momentum with respect to proper time, i.e. dp/dτ.
    In a continuous medium, the 3D density of force combines with the density of power to form a covariant 4-vector. The spatial part is the result of dividing the force on a small cell (in 3-space) by the volume of that cell. The time component is −1/c times the power transferred to that cell divided by the volume of the cell. This will be used below in the section on electromagnetism.

    Relativity and unifying electromagnetism

    Theoretical investigation in classical electromagnetism led to the discovery of wave propagation. Equations generalizing the electromagnetic effects found that finite propagation speed of the E and B fields required certain behaviors on charged particles. The general study of moving charges forms the Liénard–Wiechert potential, which is a step towards special relativity.
    The Lorentz transformation of the electric field of a moving charge into a non-moving observer's reference frame results in the appearance of a mathematical term commonly called the magnetic field. Conversely, the magnetic field generated by a moving charge disappears and becomes a purely electrostatic field in a comoving frame of reference. Maxwell's equations are thus simply an empirical fit to special relativistic effects in a classical model of the Universe. As electric and magnetic fields are reference frame dependent and thus intertwined, one speaks of electromagnetic fields. Special relativity provides the transformation rules for how an electromagnetic field in one inertial frame appears in another inertial frame.
    Maxwell's equations in the 3D form are already consistent with the physical content of special relativity, although they are easier to manipulate in a manifestly covariant form, i.e. in the language of tensor calculus.[47] See main links for more detail.

    Status

    Special relativity in its Minkowski spacetime is accurate only when the absolute value of the gravitational potential is much less than c2 in the region of interest.[48] In a strong gravitational field, one must use general relativity. General relativity becomes special relativity at the limit of weak field. At very small scales, such as at the Planck length and below, quantum effects must be taken into consideration resulting in quantum gravity. However, at macroscopic scales and in the absence of strong gravitational fields, special relativity is experimentally tested to extremely high degree of accuracy (10−20)[49] and thus accepted by the physics community. Experimental results which appear to contradict it are not reproducible and are thus widely believed to be due to experimental errors.
    Special relativity is mathematically self-consistent, and it is an organic part of all modern physical theories, most notably quantum field theory, string theory, and general relativity (in the limiting case of negligible gravitational fields).
    Newtonian mechanics mathematically follows from special relativity at small velocities (compared to the speed of light) – thus Newtonian mechanics can be considered as a special relativity of slow moving bodies. See classical mechanics for a more detailed discussion.
    Several experiments predating Einstein's 1905 paper are now interpreted as evidence for relativity. Of these it is known Einstein was aware of the Fizeau experiment before 1905,[50] and historians have concluded that Einstein was at least aware of the Michelson–Morley experiment as early as 1899 despite claims he made in his later years that it played no role in his development of the theory.[17]
    • The Fizeau experiment (1851, repeated by Michelson and Morley in 1886) measured the speed of light in moving media, with results that are consistent with relativistic addition of colinear velocities.
    • The famous Michelson–Morley experiment (1881, 1887) gave further support to the postulate that detecting an absolute reference velocity was not achievable. It should be stated here that, contrary to many alternative claims, it said little about the invariance of the speed of light with respect to the source and observer's velocity, as both source and observer were travelling together at the same velocity at all times.
    • The Trouton–Noble experiment (1903) showed that the torque on a capacitor is independent of position and inertial reference frame.
    • The Experiments of Rayleigh and Brace (1902, 1904) showed that length contraction doesn't lead to birefringence for a co-moving observer, in accordance with the relativity principle.
    Particle accelerators routinely accelerate and measure the properties of particles moving at near the speed of light, where their behavior is completely consistent with relativity theory and inconsistent with the earlier Newtonian mechanics. These machines would simply not work if they were not engineered according to relativistic principles. In addition, a considerable number of modern experiments have been conducted to test special relativity. Some examples:

    Theories of relativity and quantum mechanics

    Special relativity can be combined with quantum mechanics to form relativistic quantum mechanics. It is an unsolved problem in physics how general relativity and quantum mechanics can be unified; quantum gravity and a "theory of everything", which require such a unification, are active and ongoing areas in theoretical research.
    The early Bohr–Sommerfeld atomic model explained the fine structure of alkali metal atoms using both special relativity and the preliminary knowledge on quantum mechanics of the time.[51]
    In 1928, Paul Dirac constructed an influential relativistic wave equation, now known as the Dirac equation in his honour,[52] that is fully compatible both with special relativity and with the final version of quantum theory existing after 1926. This equation explained not only the intrinsic angular momentum of the electrons called spin, it also led to the prediction of the antiparticle of the electron (the positron),[52][53] and fine structure could only be fully explained with special relativity. It was the first foundation of relativistic quantum mechanics. In non-relativistic quantum mechanics, spin is phenomenological and cannot be explained.
    On the other hand, the existence of antiparticles leads to the conclusion that relativistic quantum mechanics is not enough for a more accurate and complete theory of particle interactions. Instead, a theory of particles interpreted as quantized fields, called quantum field theory, becomes necessary; in which particles can be created and destroyed throughout space and time.