Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Optics

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 15

OPTICS

Optics is the branch of physics which studies the behavior and properties of light, including its
interactions with matter and the construction of instruments that use or detect it.[1] Optics usually
describes the behavior 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 modeled 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 fiber optics.

Optics began with the development of lenses by the ancient Egyptians and Mesopotamians. The earliest
known lenses were made from polished crystal, often quartz, and have been dated 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 ?pt???, meaning
appearance or look.[3] Plato first articulated his emission theory, the idea that visual perception is
accomplished by rays of light emitted by the eyes and commented on the parity reversal of mirrors in
Timaeus.[4] Some hundred years later, Euclid wrote a treatise entitled Optics wherein he describes the
mathematical rules of perspective and describes the effects of refraction qualitatively.[5] Ptolemy, in his
treatise Optics, summarizes much of Euclid and goes 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.[6]
Al-Kindi (c. 801–73) was one of the earliest important writers on optics in the Islamic world. In a work
known in the West as De radiis stellarum, al-Kindi resurrected Plato's emission theory[7] which had an
influence on later Western scholars such as Robert Grosseteste and Roger Bacon.[8] In 984, the Persian
mathematician, Ibn Sahl wrote a treatise "On Burning Mirrors and Lenses", correctly describing a law of
refraction mathematically equivalent to Snell's law.[9] He used his law of refraction to compute the
shapes of lenses and mirrors that focus light at a single point on the axis. In the early 11th century,
Alhazen (Ibn al-Haytham) wrote his Book of Optics, which extensively documented the then-current
Islamic understanding of optics and revolutionized the field.[10][11][12] It included the first descriptions
of optical phenomena associated with pinholes and concave lenses,[13][14] provided the first correct
explanation of vision, described various experiments using an early scientific method,[15] and greatly
influenced the later development of the modern telescope.[16]

In the 13th century, Roger Bacon, inspired by Ibn al-Haytham, used parts of glass spheres as magnifying
glasses, and discovered that light reflects from objects rather than being released from them. In Italy,
around 1284, Salvino D'Armate invented the first wearable eyeglasses.[17] The first rudimentary
telescopes were developed independently in the 1570s and 1580s by Leonard Digges,[18] Taqi al-
Din[19] and Giambattista della Porta.[20]

The earliest known working telescopes were refracting telescopes, a type which relies entirely on lenses
for magnification. Their development in the Netherlands in 1608 was by three individuals: Hans
Lippershey and Zacharias Janssen, who were spectacle makers in Middelburg, Holland, and Jacob Metius
of Alkmaar. In Italy, Galileo greatly improved upon these designs the following year. In 1668, Isaac
Newton constructed the first practical reflecting telescope, which bears his name, the Newtonian
reflector.[21]

The first microscope was made around 1595, also in Middelburg.[22] Three different eyeglass makers
have been given credit for the invention: Lippershey (who also developed the first real telescope);
Janssen; and his father, Hans. The coining of the name "microscope" has been credited to Giovanni
Faber, who gave that name to Galileo's compound microscope in 1625.[23]

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.[24] This differed substantively from ancient Greek
notions that light emanated from the eye. In the late 1660s and early 1670s, Newton expanded
Descartes' ideas into a corpuscle theory of light, famously showing that white light, instead of being a
unique color, was really a composite of different colors that can be separated into a spectrum with a
prism. In 1690, Christian Huygens proposed a wave theory for light based on suggestions that had been
made by Robert Hooke in 1664. Hooke himself publicly criticized 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.[24]
Newtonian optics and emission theory 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, something normal particles do not follow. This work led to a theory of diffraction
for light and opened an entire area of study in physical optics.[25] Wave optics was successfully unified
with electromagnetic theory by James Clerk Maxwell in the 1860s.[26]

The next development in optical theory came in 1899 when Max Planck correctly modeled blackbody
radiation by assuming that the exchange of energy between light and matter only occurred in discrete
amounts he called quanta.[27] In 1905, Albert Einstein published the theory of the photoelectric effect
that firmly established the quantization of light itself.[28] 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.[29] 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 was the theory of quantum electrodynamics,
which explains all optics and electromagnetic processes in general as being the result of the exchange of
real and virtual photons.[30]

Quantum optics gained practical importance with the invention of the maser in 1953 and the laser in
1960.[31] 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.

Reflections
Diagram of specular reflection

Reflections can be divided into two types: specular reflection and diffuse reflection. Specular reflection
describes glossy 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 matte surfaces, 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 surface. 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.

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 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.[37] 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 magnification of a flat mirror is unity.) 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[37] retroreflect light, producing reflected rays that
travel back in the direction from which the incident rays came.

Mirrors with curved surfaces can be modeled 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.[37]

Refraction
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 n1 and another medium with
index of refraction n2. In such situations, Snell's Law describes the resulting deflection of the light ray:

where ?1 and ?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:

where v1 and v2 are the wave velocities through the respective media.[37]

Various consequences of Snell's Law include the fact that for light rays traveling 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 fiber optics technology. As light signals travel down a fiber 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 polarized 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.[37]
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.[37

Diffraction and optical resolution

Main articles: 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 ? 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'.[46][47] Later that century, Robert Hooke and Isaac Newton also described
phenomena now known to be diffraction in Newton's rings[48] while James Gregory recorded his
observations of diffraction patterns from bird feathers.[49]

The first physical optics model of diffraction that relied on Huygens' Principle was developed in 1803 by
Thomas Young in his accounts of 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.[50] In 1815 and 1818, Augustin-Jean Fresnel firmly established the mathematics of how
wave interference can account for diffraction.[39]

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 (?). In general, the equation takes the form

m? = dsin?

where d is the separation between two wavefront sources (in the case of Young's experiments, it was
two slits), ? is the angular separation between the central fringe and the mth order fringe, where the
central maximum is m = 0.[51]
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.[51] More complicated models of diffraction require working with
the mathematics of Fresnel or Fraunhofer diffraction.[40]

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.[51]

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.[39] The size of such a disk is given by

where ? is the angular resolution, ? 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.[51] Interferometry, with its ability to mimic extremely large baseline apertures, allows for
the greatest angular resolution possible.[44]

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 utilized to eliminate the atmospheric disruption of images and achieve results that
approach the diffraction limit.[52]

Modern optics

Main articles: Optical physics and Optical engineering

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.[59]

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.[60]

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.[61] A professional community of
researchers in nonlinear optics has developed in the last several decades due to advances in laser
technology.[62]

Optical instruments

Illustrations of various optical instruments from the 1728 Cyclopaedia

Main article: Optical instruments

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.[77]

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.[77] 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.[78]

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.[77]

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.[77]

A wave is a disturbance that propagates through space and time, usually with transference of energy. A
mechanical wave is a wave that propagates or travels through a medium due to the restoring forces it
produces upon deformation. For example, when a sound wave is traveling through the air, air molecules
slam into their neighbors, which pushes their neighbors into their neighbors (and so on); but when air
molecules collide with their neighbors, they also bounce away from them back in the direction they
came from. These collisions provide a restoring force that keeps the molecules from actually traveling
with the wave.

Waves travel and transfer energy from one point to another, often with no permanent displacement of
the particles of the medium—that is, with little or no associated mass transport; they consist instead of
oscillations or vibrations around almost fixed locations. In the picture of water waves, if we imagine a
cork on the water, it would bob up and down staying in about the same place, although the wave itself is
moving outward. When we say that a wave carries energy but not mass, we are referring to this fact that
even as the wave travels outward from the center (carrying energy of motion), the medium itself does
not flow with it.

In many areas of science, the idea of a wave is used metaphorically. If an ocean wave is seen as a
prototype wave, it is the basis for the metaphor—the surface of water undulating up and down.
However, upon investigating a sound wave, its air does not undulate up and down (as the ocean surface
did). Instead, an abstraction is made; if we could look at the air molecules, they would be bunching
together (in compressions) and then spreading apart (in rarefactions). Thus, the medium itself is not
undulating up and down, but its density is (and its pressure is). When we speak of waves in physics,
therefore, we are often speaking metaphorically, in an abstraction, of a periodic fluctuation of a specific
characteristic. The characteristics that oscillate could be density, pressure, electrical or magnetic
polarities or other (sometimes exotic) characteristics.

There are also waves capable of traveling through a vacuum, including electromagnetic radiation.
Ultraviolet radiation, infrared radiation, gamma rays, X-rays, and radio waves are examples of these
types of waves. They consist of period oscillations in electrical and magnetic properties that grow, reach
a peak, diminish, go to zero, and then continue these changes in a periodic fashion.

As well, it is believed that gravitational waves travel through space; gravitational waves have never been
directly detected but are believed to exist.

Agreeing on a single, all-encompassing definition for the term wave is non-trivial. A vibration can be
defined as a back-and-forth motion around a reference value. However, a vibration is not necessarily a
wave. Defining the necessary and sufficient characteristics that qualify a phenomenon to be called a
wave is, at least, flexible.

The term is often understood intuitively as the transport of disturbances in space, not associated with
motion of the medium occupying this space as a whole. In a wave, the energy of a vibration is moving
away from the source in the form of a disturbance within the surrounding medium (Hall 1980, p. 8).
However, this notion is problematic for a standing wave (for example, a wave on a string), where energy
is moving in both directions equally, or for electromagnetic / light waves in a vacuum, where the
concept of medium does not apply. There are water waves in the ocean; light waves from the sun;
microwaves inside the microwave oven; radio waves transmitted to the radio; and sound waves from
the radio, telephone, and voices.

It may be seen that the description of waves is accompanied by a heavy reliance on physical origin when
describing any specific instance of a wave process. For example, acoustics is distinguished from optics in
that sound waves are related to a mechanical rather than an electromagnetic wave-like transfer /
transformation of vibratory energy. Concepts such as mass, momentum, inertia, or elasticity, become
therefore crucial in describing acoustic (as distinct from optic) wave processes. This difference in origin
introduces certain wave characteristics particular to the properties of the medium involved (for
example, in the case of air: vortices, radiation pressure, shock waves, etc., in the case of solids: Rayleigh
waves, dispersion, etc., and so on).

Other properties, however, although they are usually described in an origin-specific manner, may be
generalized to all waves. For such reasons, wave theory represents a particular branch of physics that is
concerned with the properties of wave processes independently from their physical origin.[1] For
example, based on the mechanical origin of acoustic waves there can be a moving disturbance in space–
time if and only if the medium involved is neither infinitely stiff nor infinitely pliable. If all the parts
making up a medium were rigidly bound, then they would all vibrate as one, with no delay in the
transmission of the vibration and therefore no wave motion. On the other hand, if all the parts were
independent, then there would not be any transmission of the vibration and again, no wave motion.
Although the above statements are meaningless in the case of waves that do not require a medium,
they reveal a characteristic that is relevant to all waves regardless of origin: within a wave, the phase of
a vibration (that is, its position within the vibration cycle) is different for adjacent points in space
because the vibration reaches these points at different times.

Similarly, wave processes revealed from the study of waves other than sound waves can be significant to
the understanding of sound phenomena. A relevant example is Thomas Young's principle of interference
(Young, 1802, in Hunt 1992, p. 132). This principle was first introduced in Young's study of light and,
within some specific contexts (for example, scattering of sound by sound), is still a researched area in
the study of sound.

Characteristics

Periodic waves are characterized by crests (highs) and troughs (lows), and may usually be categorized as
either longitudinal or transverse. Transverse waves are those with vibrations perpendicular to the
direction of the propagation of the wave; examples include waves on a string, and electromagnetic
waves. Longitudinal waves are those with vibrations parallel to the direction of the propagation of the
wave; examples include most sound waves.

When an object bobs up and down on a ripple in a pond, it experiences an orbital trajectory because
ripples are not simple transverse sinusoidal waves.

A = In deep water.
B = In shallow water. The elliptical movement of a surface particle becomes flatter with decreasing
depth.

1 = Progression of wave

2 = Crest

3 = Trough

Ripples on the surface of a pond are actually a combination of transverse and longitudinal waves;
therefore, the points on the surface follow orbital paths.

All waves have common behavior under a number of standard situations. All waves can experience the
following:

• Reflection — change in wave direction after it strikes a reflective surface, causing the angle the
wave makes with the reflective surface in relation to a normal line to the surface to equal the angle the
reflected wave makes with the same normal line

• Refraction — change in wave direction because of a change in the wave's speed from entering a
new medium

• Diffraction — bending of waves as they interact with obstacles in their path, which is more
pronounced for wavelengths on the order of the diffracting object size

• Interference — superposition of two waves that come into contact with each other (collide)

• Dispersion — wave splitting up by frequency

• Rectilinear propagation — the movement of light waves in a straight line also helpful for
seismographs

[edit] Polarization

Main article: Polarization (waves)

A wave is polarized if it oscillates in one direction or plane. A wave can be polarized by the use of a
polarizing filter. The polarization of a transverse wave describes the direction of oscillation in the plane
perpendicular to the direction of travel.

Longitudinal waves such as sound waves do not exhibit polarization. For these waves the direction of
oscillation is along the direction of travel.this is very important

[edit] Examples
An ocean surface wave crashing into rocks

Examples of waves include:

• Ocean surface waves, which are perturbations that propagate through water

• Radio waves, microwaves, infrared rays, visible light, ultraviolet rays, x-rays, and gamma rays,
which make up electromagnetic radiation; can be propagated without a medium, through vacuum; and
travel at 299,792,458 m/s in a vacuum

• Sound — a mechanical wave that propagates through gases, liquids, solids and plasmas

• Waves of traffic, that is, propagation of different densities of motor vehicles, and so forth, which
can be modeled as kinematic waves, as first presented by Sir M. J. Lighthill[2]

• Seismic waves in earthquakes, of which there are three types, called S, P, and L

• Gravitational waves, which are nonlinear fluctuations in the curvature of spacetime predicted by
General Relativity, but which have yet to be observed empirically

• Inertial waves, which occur in rotating fluids and are restored by the Coriolis effect

• Matter waves, which describe the fundamental wave nature of matter.

[edit] Mathematical description

Wavelength of a cosine wave, ?, can be measured between any two points with the same phase, such as
between crests, or troughs, or corresponding zero crossings as shown.

[] Sinusoidal waves

Mathematically, the most basic wave is the sine wave (or harmonic wave or sinusoid), with an amplitude
u described by the equation:

where A is the semi-amplitude of the wave, half the peak-to-peak amplitude, often called simply the
amplitude – the maximum distance from the highest point of the disturbance in the medium (the crest)
to the equilibrium point during one wave cycle. In the illustration to the right, this is the maximum
vertical distance between the baseline and the wave; x is the space coordinate, t is the time coordinate,
k is the wavenumber (spatial frequency), ? is the temporal frequency, and f is a phase offset.
The units of the semi-amplitude depend on the type of wave — waves on a string have an amplitude
expressed as a distance (meters), sound waves as pressure (pascals) and electromagnetic waves as the
amplitude of the electric field (volts/meter).

The wavelength (denoted as ?) is the distance between two sequential crests (or troughs), and generally
is measured in meters.

A wavenumber k, the spatial frequency of the wave in radians per unit distance (typically per meter),
can be associated with the wavelength by the relation

Sine waves correspond to simple harmonic motion.

The period T is the time for one complete cycle of an oscillation of a wave. The frequency f (also
frequently denoted as ? ) is the number of periods per unit time (per second) and is measured in hertz.
These are related by:

In other words, the frequency and period of a wave are reciprocals.

The angular frequency ? represents the frequency in radians per second. It is related to the frequency by

Various local wavelengths on a crest-to-crest basis in an ocean wave approaching shore.[3]

The wavelength ? of a sinusoidal waveform traveling at constant speed v is given by:[4]

Refraction: when a plane wave encounters a medium in which it has a slower speed, the wavelength
decreases, and the direction adjusts accordingly.

where v is called the phase speed (magnitude of the phase velocity) of the wave and f is the wave's
frequency.
Wavelength can be a useful concept even if the wave is not periodic in space. For example, in an ocean
wave approaching shore, shown in the figure, the incoming wave undulates with a varying local
wavelength that depends in part on the depth of the sea floor compared to the wave height. The
analysis of the wave can be based upon comparison of the local wavelength with the local water depth.
[3]

Although arbitrary wave shapes will propagate unchanged in lossless linear time-invariant systems, in
the presence of dispersion the sine wave is the unique shape that will propagate unchanged but for
phase and amplitude, making it easy to analyze.[5] Due to the Kramers–Kronig relations, a linear
medium with dispersion also exhibits loss, so the sine wave propagating in a dispersive medium is
attenuated in certain frequency ranges that depend upon the medium.[6] The sine function is periodic,
so the sine wave or sinusoid has a wavelength in space and a period in time.[7][8]

The sinusoid is defined for all times and distances, whereas in physical situations we usually deal with
waves that exist for a limited span in space and duration in time. Fortunately, an arbitrary wave shape
can be decomposed into an infinite set of sinusoidal waves by the use of Fourier analysis. As a result, the
simple case of a single sinusoidal wave can be applied to more general cases.[9][10] In particular, many
media are linear, or nearly so, so the calculation of arbitrary wave behavior can be found by adding up
responses to individual sinusoidal waves using the superposition principle to find the solution for a
general waveform.[11] When a medium is nonlinear, the response to complex waves cannot be
determined from a sine-wave decomposition.

[edit] The wave equation

Main articles: Wave equation and D'Alembert's formula

See also: Telegrapher's equations

The wave equation is a partial differential equation that describes the evolution of a wave over time in a
medium where the wave propagates at the same speed independent of wavelength (no dispersion), and
independent of amplitude (linear media, not nonlinear).[12] General solutions are based upon
Duhamel's principle.[13]

In particular, consider the wave equation in one dimension, for example, as applied to a string. Suppose
a one-dimensional wave is traveling along the x axis with velocity v and amplitude u (which generally
depends on both x and t), the wave equation is

The velocity v will depend on the medium through which the wave is moving.

The general solution for the wave equation in one dimension was given by d'Alembert; it is known as
d'Alembert's formula:[14]
This formula represents two shapes traveling through the medium in opposite directions; F in the
positive x direction, and G in the negative x direction, of arbitrary functional shapes F and G.

[edit] Spatial and temporal relationships

See also: Wavelength

Wavelength of an irregular periodic waveform at a particular moment in time based upon the crest-to-
crest or trough-to-trough definition of ?.[15]

The form of the forward propagating wave F in d'Alembert's formula involves the argument x - vt.
Constant values of this argument correspond to constant values of F, and these constant values occur if
x increases at the same rate that vt increases. That is, the wave shaped like the function F will move in
the positive x-direction at velocity v (and G will propagate at the same speed in the negative x-direction).
[16]

In the case of a periodic function F with period ?, that is, F(x + ? - vt) = F(x - vt), the periodicity of F in
space means that a snapshot of the wave at a given time t finds the wave varying periodically in space
with period ? (sometimes called the wavelength of the wave). In a similar fashion, this periodicity of F
implies a periodicity in time as well: F(x - v(t + T)) = F(x - vt) provided vT = ?, so an observation of the
wave at a fixed location x finds the wave undulating periodically in time with period T = ?/v.[15]

To summarize:

"A function F (x) is periodic if F(x+?) = F(x), for all x. The constant ? is called a period of the function. The
smallest such period is called the fundamental period or simply the period of F. If x represents a space
coordinate, then the period may instead be called the wavelength and is often written ?; if it represents
the time coordinate, the period might instead be denoted by T." Flowers, p. 473[17]

You might also like