Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 27

Photochemistry, a sub-discipline of chemistry, is the study of the interactions

between light and atoms or molecules.[1] Photochemistry describes chemical


reactions that proceed with the absorption of light. Everyday examples include
photosynthesis, the degradation of plastics and the formation of vitamin D with
sunlight.

Principles
Light is a type of electromagnetic radiation, a source of energy. The GrotthussDraper law (for
chemists Theodor Grotthuss and John W. Draper), states that light must be absorbed by a
chemical substance in order for a photochemical reaction to take place.
The second law of photochemistry, the Stark-Einstein law, states that for each photon of light
absorbed by a chemical system, only one molecule is activated for a photochemical reaction.
This law, also known as the photoequivalence law, was derived by Albert Einstein at the time
when the quantum (photon) theory of light was being developed.
Chemical reactions occur only when a molecule is provided the necessary "activation energy". A
simple example can be the combustion of gasoline (a hydrocarbon) into carbon dioxide and
water. In this reaction, the activation energy is provided in the form of heat or a spark. In case of
photochemical reactions light provides the activation energy. Simplistically, light is one
mechanism for providing the activation energy required for many reactions. If laser light is
employed, it is possible to selectively excite a molecule so as to produced a desired electronic
and vibrational state. Equally, the emission from a particular state may be selectively monitored,
providing a measure of the population of that state. If the chemical system is at low pressure, this
enables scientists to observe the energy distribution of the products of a chemical reaction before
the differences in energy have been smeared out and averaged by repeated collisions.
The absorption of a photon of light by a reactant molecule may also permit a reaction to occur
not just by bringing the molecule to the necessary activation energy, but also by changing the
symmetry of the molecule's electronic configuration, enabling an otherwise inaccessible reaction
path, as described by the Woodward-Hoffmann selection rules. A 2+2 cycloaddition reaction is
one example of a pericyclic reaction that can be analyzed using these rules or by the related
frontier molecular orbital theory.
Photochemical reactions involve electronic reorganization initiated by electromagnetic radiation.
The reactions are several orders of magnitude faster than thermal reactions; reactions as fast as
109 seconds and associated processes as fast as 1015 seconds are often observed.

Spectral regions
Photochemists typically work in only a few sections of the electromagnetic spectrum. Some of
the most widely used sections, and their wavelengths, are the following:

Ultraviolet: 100400 nm
Visible Light: 400700 nm

Near infrared: 7002500 nm

Applications
Many important processes involve photochemistry. The premier example is photosynthesis, in
which most plants use solar energy to convert carbon dioxide and water into glucose, disposing
of oxygen as a side-product. Humans rely on photochemistry for the formation of vitamin D. In
fireflies, an enzyme in the abdomen catalyzes a reaction that results in bioluminescence.[2]
Photochemistry can also be highly destructive. Medicine bottles are often made with darkened
glass to prevent the drugs from photodegradation. A pervasive reaction is the generation of
singlet oxygen by photosensitized reactions of triplet oxygen. Typical photosensitizers include
tetraphenylporphyrin and methylene blue. The resulting singlet oxygen is an aggressive oxidant,
capable of converting C-H bonds into C-OH groups.In photodynamic therapy, light is used to
destroy tumors by the action of singlet oxygen.
Many polymerizations are started by photoinitiatiors, which decompose upon absorbing light to
produce the free radicals for Radical polymerization.

In the area of photochemistry, a photochemical reaction is a chemical reaction that is induced


by light. Photochemical reactions are valuable in organic and inorganic chemistry because they
proceed differently than thermal reactions. Photochemical reactions are not only very useful but
also can be a serious nuisance, as in the photodegradation of many materials, e.g. polyvinyl
chloride. A large-scale application of photochemistry is photoresist technology, used in the
production of microelectronic components. Vision is initiated by a photochemical reaction of
rhodopsin.[3]

Experimental set-up
Photochemical reactions require a light source that emits wavelengths corresponding to an
electronic transition in the reactant. In the early experiments (and in everyday life), sunlight was
the light source, although it is polychromatic. Mercury-vapor lamps are more common in the
laboratory. Low pressure mercury vapor lamps mainly emit at 254 nm. For polychromatic
sources, wavelength ranges can be selected using filters. Alternatively, LEDs and Rayonet lamps
emit monochromatically.

Schlenk tube containing slurry of orange crystals of Fe2(CO)9 in acetic acid after its
photochemical synthesis from Fe(CO)5. The mercury lamp (connected to white power cords) can
be seen on the left, set inside a water-jacketed quartz tube.
The emitted light must of course reach the targeted functional group without being blocked by
the reactor, medium, or other functional groups present. For many applications, quartz is used for
the reactors as well as to contain the lamp. Pyrex absorbs at wavelengths shorter than 275 nm.
The solvent is an important experimental parameter. Solvents are potential reactants and for this
reason, chlorinated solvents are avoided because the C-Cl bond can lead to chlorination of the
substrate. Strongly absorbing solvents prevent photons from reaching the substrate. Hydrocarbon
solvents absorb only at short wavelengths and are thus preferred for photochemical experiments
requiring high energy photons. Solvents containing unsaturation absorb at longer wavelengths
and can usefully filter out short wavelengths. For example, cyclohexane and acetone "cut off"
(absorb strongly) at wavelengths shorter than 215 and 330 nm, respectively.

Excitation
Photoexcitation is the first step in a photochemical process where the reactant is elevated to a
state of higher energy, an excited state. The photon can be absorbed directly by the reactant or by
a photosensitizers, which absorbs the photon and transfers the energy to the reactant. The
opposite process is called quenching when a photoexited state is deactivated by a chemical
reagent.
Most photochemical transformations occur through a series of simple steps known as primary
photochemical processes. One common example of these processes is the excited state proton
transfer (ESPT).

Organic photochemistry
Examples of photochemical organic reactions are electrocyclic reactions, photoisomerization and
Norrish reactions.
Alkenes undergo many important reactions that proceed via a photon-induced to * transition.
The first electronic excited state of an alkene lack the -bond, so that rotation about the C-C
bond is rapid and the molecule engages in reactions not observed thermally. These reactions
include cis-trans isomerization, cycloaddition to other (ground state) alkene to give cyclobutane
derivatives. The cis-trans isomerization of a (poly)alkene is involved in retinal, a component of
the machinery of vision. The dimerization of alkenes is relevant to the photodamage of DNA,
where thymine dimers are observed upon illuminating DNA to UV radiation. Such dimers
interfere with transcription. The beneficial effects of sunlight are associated with the
photochemically induced retro-cyclization (decyclization) reaction of ergosterol to give vitamin
D. In the DeMayo reaction, an alkene reacts with a 1,3-diketone reacts via its enol to yield a 1,5diketone. Still another common photochemical reaction is Zimmerman's Di-pi-methane
rearrangement.
In an industrial application, about 100,000 tonnes of benzyl chloride are prepared annually by the
gas-phase photochemical reaction of toluene with chlorine.[4] The light is absorbed by chlorine
molecule, the low energy of this transition being indicted by the yellowish color of the gas. The
photon induces homolysis of the Cl-Cl bond, and the resulting chlorine radical converts toluene
to the benzyl radical:
Cl2 + h 2 Cl
C6H5CH3 + Cl C6H5CH2 + HCl
C6H5CH2 + Cl C6H5CH2Cl
Mercaptans can be produced by photochemical addition of hydrogen sulfide (H2S) to alpha
olefins.

Inorganic and organometallic photochemistry


Coordination complexes and organometallic compounds are also photoreactive. These reactions
can entail cis-trans isomerization. More commonly photoreactions result in dissociation of
ligands, since the photon excites an electron on the metal to an orbital that is antibonding with
respect to the ligands. Thus, metal carbonyls that resist thermal substitution undergo
decarbonylation upon irradiation with UV light. UV-irradiation of a THF solution of
molybdenum hexacarbonyl gives the THF complex, which is synthetically useful:
Mo(CO)6 + THF Mo(CO)5(THF) + CO
In a related reaction, photolysis of iron pentacarbonyl affords diiron nonacarbonyl (see figure):
2 Fe(CO)5 Fe2(CO)9 + CO

History
Although bleaching has long been practiced, the first photochemical reaction was described by
Trommsdorf in 1834.[5] He observed that crystals of the compound -santonin when exposed to
sunlight turned yellow and burst. In a 2007 study the reaction was described as a succession of
three steps taking place within a single crystal.[6]

The first step is a rearrangement reaction to a cyclopentadienone intermediate 2, the second one a
dimerization in a Diels-Alder reaction (3) and the third one a intramolecular [2+2]cycloaddition
(4). The bursting effect is attributed to a large change in crystal volume on dimerization.

References
1.
2.

^ IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online
corrected version: (2006-) "photochemistry".
^ David Stanley Saunders Insect clocks, Elsevier, 2002, ISBN 0444504079 p. 179

3.

^ Christophe Dugave Cis-trans isomerization in biochemistry, Wiley-VCH, 2006 ISBN


3527313044 p. 56

4.

^ M. Rossberg et al. Chlorinated Hydrocarbons in Ullmanns Encyclopedia of Industrial


Chemistry 2006, Wiley-VCH, Weinheim. doi:10.1002/14356007.a06_233.pub2

5.
6.

^ Trommsdorf, Ann. Chem. Pharm. 1834, 11


^ The Photoarrangement of -Santonin is a Single-Crystal-to-Single-Crystal Reaction: A Long
Kept Secret in Solid-State Organic Chemistry Revealed Arunkumar Natarajan, C. K. Tsai, Saeed I. Khan,

Patrick McCarren, K. N. Houk, and Miguel A. Garcia-Garibay J. Am. Chem. Soc., 129 (32), 9846 -9847,
2007. doi:10.1021/ja073189o

See also

Journal of Photochemistry and Photobiology


Photoelectrochemical cell

Photochemical and Photobiological Sciences

Photochemistry and Photobiology

Photochemical Logic Gates

Photosynthesis

Photoelectrochemical processes usually involve transforming light into other


forms of energy.[1] These processes apply to photochemistry, optically pumped
lasers, sensitized solar cells, luminescence, and the effect of reversible change of
color upon exposure to light. To the right photons are emitted in a coherent beam
from a laser

Electron excitation

After absorbing energy, an electron may jump from the ground state to a higher
energy excited state.

Electron excitation is the movement of an electron to a higher energy state. This can either be
done by photoexcitation (PE), where the original electron absorbs the photon and gains all the
photon's energy or by electrical excitation (EE), where the original electron absorbs the energy of
another, energetic electron. Within a semiconductor crystal lattice, thermal excitation is a process
where lattice vibrations provide enough energy to move electrons to a higher energy band. When

an excited electron falls back to a lower energy state again, it is called electron relaxation. This
can be done by radiation of a photon or giving the energy to a third spectator particle as well.[2]
In physics there is a specific technical definition for energy level which is often associated with
an atom being excited to an excited state. The excited state, in general, is in relation to the
ground state, where the excited state is at a higher energy level than the ground state.

Photoexcitation
Photoexcitation is the mechanism of electron excitation by photon absorption, when the energy
of the photon is too low to cause photoionization. The absorption of photon takes place in
accordance to the Planck's Quantum Theory.
Photoexcitation plays role in photoisomerization. Photoexcitation is exploited in dye-sensitized
solar cells, photochemistry, luminescence, optically pumped lasers, and in some photochromic
applications.
See also: Photoelectric effect

Photoisomerization
In chemistry, photoisomerization is molecular behavior in which structural change between
isomers is caused by photoexcitation. Both reversible and irreversible photoisomerization
reactions exist. However, the word "photoisomerization" usually indicates a reversible process.
Photoisomerizable molecules are already put to practical use, for instance, in pigments for
rewritable CDs, DVDs, and 3D optical data storage solutions. In addition, recent interest in
photoisomerizable molecules has been aimed at molecular devices, such as molecular switches,
molecular motors, and molecular electronics.
Photoisomerization behavior can be roughly categorized into several classes: trans (or E) and cis
(or Z) conversion, and open ring and closed ring transition. Instances of the former include
stilbene and azobenzene. This class of compounds has a double bond, and rotation or inversion
around the double bond affords isomerization between the two states. Examples of the latter
include fulgide and diarylethene. These types of compounds undergo bond cleavage and bond
creation upon irradiation with particular wavelengths of light. Sill another type is the Di-pimethane rearrangement.

Photoionization
Photoionization is the physical process in which an incident photon ejects one or more electrons
from an atom, ion or molecule. This is essentially the same process that occurs with the
photoelectric effect with metals. In the case of a gas, the term photoionization is more common.[3]
The ejected electrons, known as photoelectrons, carry information about their pre-ionized states.
For example, a single electron can have a kinetic energy equal to the energy of the incident

photon minus the electron binding energy of the state it left. Photons with energies less than the
electron binding energy may be absorbed or scattered but will not photoionize the atom or ion.[3]
For example, to ionize hydrogen, photons need an energy greater than 13.6 electronvolts, which
corresponds to a wavelength of 91.2 nm.[4] For photons with greater energy than this, the energy
of the emitted photoelectron is given by:

where h is Planck's constant and is the frequency of the photon.


This formula defines the photoelectric effect.
Not every photon which encounters an atom or ion will photoionize it. The probability of
photoionization is related to the photoionization cross-section, which depends on the energy of
the photon and the target being considered. For photon energies below the ionization threshold,
the photoionization cross-section is near zero. But with the development of pulsed lasers it has
become possible to create extremely intense, coherent light where multi-photon ionization may
occur. At even higher intensities (around 1015 - 1016 W/cm2 of infrared or visible light), nonperturbative phenomena such as barrier suppression ionization[5] and rescattering ionization[6]
are observed.
[edit] Multi-photon ionization

Several photons of energy below the ionization threshold may actually combine their energies to
ionize an atom. This probability decreases rapidly with the number of photons required, but the
development of very intense, pulsed lasers still makes it possible. In the perturbative regime
(below about 1014 W/cm2 at optical frequencies), the probability of absorbing N photons depends
on the laser-light intensity I as IN .[7]
Above-threshold ionization (ATI) [8] is an extension of multi-photon ionization where even more
photons are absorbed than actually would be necessary to ionize the atom. The excess energy
gives the released electron higher kinetic energy than the usual case of just-above threshold
ionization. More precisely, The released electron will have an integer number of photon-energies
more kinetic energy than in the normal (lowest possible number of photons) ionization.
See also
Main articles: Fluorescence spectroscopy, Fluorescence, and Photoionization
mode

Photo-Dember
Main article: Photo-Dember

In semiconductor physics the Photo-Dember effect (named after its discoverer H. Dember)
consists in the formation of a charge dipole in the vicinity of a semiconductor surface after ultrafast photo-generation of charge carriers. The dipole forms owing to the difference of mobilities
(or diffusion constants) for holes and electrons which combined with the break of symmetry
provided by the surface lead to an effective charge separation in the direction perpendicular to
the surface.[9]

GrotthussDraper law
The GrotthussDraper law (also called Principle of Photochemical Activation) states that only
that light which is absorbed by a system can bring about a photochemical change. Materials such
as dyes and phosphors must be able to absorb "light" at optical frequencies. A basis for
Fluorescence and phosphorescence is found in this law. It was first proposed in 1817 by Theodor
Grotthuss and John W. Draper. This is considered to be one of the two basic laws of
photochemistry. The second law is the StarkEinstein law, which says that primary chemical or
physical reactions occur with each photon absorbed.[10]

StarkEinstein law
The StarkEinstein law is named after the German-born physicists Johannes Stark and Albert
Einstein, who independently formulated the law between 1908 and 1913. It is known also as the
photochemical equivalence law or photoequivalence law. In essence it says that every photon
that is absorbed will cause a (primary) chemical or physical reaction.[11]
The photon is a quantum of radiation, or one unit of radiation. Therefore, this is a single unit of
EM radiation that is equal to Planck's constant (h) times the frequency of light. This quantity is
symbolized by
The photochemical equivalence law is also restated as follows: for every mole of a substance that
reacts, an equivalent mole of quanta of light are absorbed. The formula is:[11]

Emol = NAh
where NA is Avogadro's number.
The photochemical equivalence law applies to the part of a light-induced reaction that is referred
to as the primary process (i.e. absorption or fluorescence).[11]
In most photochemical reactions the primary process is usually followed by so-called secondary
photochemical processes that are normal interactions between reactants not requiring absorption
of light. As a result such reactions do not appear to obey the one quantumone molecule reactant
relationship.[11]
The law is further restricted to conventional photochemical processes using light sources with
moderate intensities; high-intensity light sources such as those used in flash photolysis and in

laser experiments are known to cause so-called biphotonic processes; i.e., the absorption by a
molecule of a substance of two photons of light.[11]

Absorption (electromagnetic radiation)


Main article: Absorption (electromagnetic radiation)
In physics, absorption of electromagnetic radiation is the way by which the energy of a photon
is taken up by matter, typically the electrons of an atom. Thus, the electromagnetic energy is
transformed to other forms of energy, for example, to heat. The absorption of light during wave
propagation is often called attenuation. Usually, the absorption of waves does not depend on their
intensity (linear absorption), although in certain conditions (usually, in optics), the medium
changes its transparency dependently on the intensity of waves going through, and the Saturable
absorption (or nonlinear absorption) occurs.

Photosensitization
Photosensitization is a process of transferring the energy of absorbed light. After absorption, the
energy is transferred to the (chosen) reactants. This is part of the work of photochemistry in
general. In particular this process is commonly employed where reactions require light sources of
certain wavelengths that are not readily available.[12]
For example, mercury absorbs radiation at 1849 and 2537 angstroms, and the source is often
high-intensity mercury lamps. It is a commonly used sensitizer. When mercury vapor is mixed
with ethylene, and the compound is irradiated with a mercury lamp, this results in the
photodecomposition of ethylene to acetylene. This occurs on absorption of light to yield excited
state mercury atoms, which are able to transfer this energy to the ethylene molecules, and are in
turn deactivated to their initial energy state.[12]
Cadmium; some of the noble gases, for example (usually) xenon; zinc; benzophenone; and a
large number of organic dyes, are also used as sensitizers.[12]
Photosensitisers are a key component of photodynamic therapy used to treat cancers.

Sensitizer
"Sensitizer" redirects here. For the particulate material used to create voids that aid in the
initiation or propagation of an explosive's detonation wave, see Explosive sensitiser.
A sensitizer in chemoluminescence is a chemical compound, capable of light emission after it
has received energy from a molecule, which became excited previously in the chemical reaction.
A good example is this:
When an alkaline solution of sodium hypochlorite and a concentrated solution of hydrogen
peroxide are mixed, a reaction occurs:

ClO-(aq) + H2O2(aq) O2*(g) + H+(aq) + Cl-(aq) + OH-(aq)


O2*is excited oxygen - meaning, one or more electrons in the O2 molecule have been promoted
to higher-energy molecular orbitals. Hence, oxygen produced by this chemical reaction somehow
'absorbed' the energy released by the reaction and became excited. This energy state is unstable,
therefore it will return to the ground state by lowering its energy. It can do that in more than one
way:

it can react further, without any light emission


it can lose energy without emission, for example, giving off heat to the surroundings or
transferring energy to another molecule

it can emit light

The intensity, duration and color of emitted light depend on quantum and kinetical factors.
However, excited molecules are frequently less capable of light emission in terms of brightness
and duration when compared to sensitizers. This is because sensitizers can store energy (that is,
be excited) for longer periods of time than other excited molecules. The energy is stored through
means of quantum vibration, so sensitizers are usually compounds which either include systems
of aromatic rings or many conjugated double and triple bonds in their structure. Hence, if an
excited molecule transfers its energy to a sensitizer thus exciting it, longer and easier to quantify
light emission is often observed.
The color (that is, the wavelength), brightness and duration of emission depend upon the
sensitizer used. Usually, for a certain chemical reaction, many different sensitizers can be used.

List of some common sensitizers

Violanthrone
Isoviolanthrone

Fluoresceine

Rubrene

9,10-diphenylanthracene

Tetracene

13,13'-dibenzantronile

Levulinic Acid

Fluorescence spectroscopy
Main articles: Fluorescence spectroscopy and Fluorescence

Fluorescence spectroscopy aka fluorometry or spectrofluorometry, is a type of electromagnetic


spectroscopy which analyzes fluorescence from a sample. It involves using a beam of light,
usually ultraviolet light, that excites the electrons in molecules of certain compounds and causes
them to emit light of a lower energy, typically, but not necessarily, visible light. A
complementary technique is absorption spectroscopy.[13][14]
Devices that measure fluorescence are called fluorometers or fluorimeters.

Absorption spectroscopy
Main article: Absorption spectroscopy
Absorption spectroscopy refers to spectroscopic techniques that measure the absorption of
radiation, as a function of frequency or wavelength, due to its interaction with a sample. The
sample absorbs energy, i.e., photons, from the radiating field. The intensity of the absorption
varies as a function of frequency, and this variation is the absorption spectrum. Absorption
spectroscopy is performed across the electromagnetic spectrum.[13][14]

See also

Electron binding energy


Isomerization

Photoionization mode

Photochromism

Photoelectric effect

Photoionization detector

References
1.

2.

^ Schiavello, Mario; NATO (1985-02). Photoelectrochemistry, Photocatalysis and Photoreactors


Fundamentals and Developments. Springer London, Limited. pp. 39. ISBN 9789027719461.
http://books.google.com/?
id=rLRMeP1KGhsC&pg=PR9&dq=Photoelectrochemical+processes#v=onepage&q=Photoelectrochemica
l%20processes&f=false.
^ Madden, R.P.; Codling, K (1965-02). "Two electron states in Helium". Astrophysical Journal
141: 364. Bibcode 1965ApJ...141..364M. doi:10.1086/148132

3.

^ a b "Radiation". Encyclopdia Britannica Online. Photoelectric. effect. 2009. pp. 1.


http://www.britannica.com/EBchecked/topic/488507/radiation. Retrieved 2009-11-09.

4.

^ Carroll, B. W.; Ostlie, D. A. (2007). An Introduction to Modern Astrophysics. London: AddisonWesley. p. 121. ISBN 0321442849.

5.

^ http://www.iop.org/EJ/abstract/1063-7869/41/5/R03

6.

^ http://ieeexplore.ieee.org/stamp/stamp.jsp?arnumber=01549346

7.

^ Deng, Z; Eberly, J H (March 1985). "Multiphoton absorption above ionization threshold by


atoms in strong laser fields". J. Opt. Soc. Am. B 2 (3): 491.
http://prola.aps.org/abstract/PRL/v42/i17/p1127_1.

8.

^ Agostini, P; Fabre, F; Mainfray, G; Petite, G; Rahman, N K (23 April 1979). "Free-Free


Transitions Following Six-Photon Ionization of Xenon Atoms". Phys. Rev. Lett. 42 (17): 11271130.
doi:10.1103/PhysRevLett.42.1127. http://prola.aps.org/abstract/PRL/v42/i17/p1127_1.

9.

^ Dekorsy, T.; Auer, H.; Bakker, H. J.; Roskos, H. G.; Kurz, H. (1996). "THz electromagnetic
emission by coherent infrared-active phonons". Physical Review B 53 (7): 4005.
doi:10.1103/PhysRevB.53.4005.

10.

^ "Radiation". Encyclopdia Britannica Online. radiation. (physics): Photochemistry. 2009.


pp. 1. http://www.britannica.com/EBchecked/topic/488507/radiation. Retrieved 2009-11-09.

11.

^ a b c d e "Photoequivalence law". Encyclopdia Britannica Online. 2009-11

12.

^ a b c "Photosensitization". Encyclopdia Britannica. 2009.. Online. 2009. pp. 1.


http://www.britannica.com/EBchecked/topic/458153/photosensitization. Retrieved 2009-11-10.

13.
14.

^ a b Modern Spectroscopy (Paperback) by J. Michael Hollas ISBN 0470844167


^ a b Symmetry and Spectroscopy: An Introduction to Vibrational and Electronic Spectroscopy
(Paperback) by Daniel C. Harris, Michael D. Bertolucci ISBN 048666144X

Electromagnetic radiation (often abbreviated E-M radiation or EMR) is a form of energy


exhibiting wave-like behavior as it travels through space. EMR has both electric and magnetic
field components, which oscillate in phase perpendicular to each other and perpendicular to the
direction of energy propagation.
Electromagnetic radiation is classified according to the frequency of its wave. In order of
increasing frequency and decreasing wavelength, these are radio waves, microwaves, infrared
radiation, visible light, ultraviolet radiation, X-rays and gamma rays (see Electromagnetic
spectrum). The eyes of various organisms sense a small and somewhat variable window of
frequencies called the visible spectrum. The photon is the quantum of the electromagnetic
interaction and the basic "unit" of light and all other forms of electromagnetic radiation and is
also the force carrier for the electromagnetic force.
EM radiation carries energy and momentum that may be imparted to matter with which it
interacts.
Theory

Shows the relative wavelengths of the electromagnetic waves of three different


colors of light (blue, green and red) with a distance scale in micrometres along the
x-axis.
Main article: Maxwell's equations

James Clerk Maxwell first formally postulated electromagnetic waves. These were subsequently
confirmed by Heinrich Hertz. Maxwell derived a wave form of the electric and magnetic
equations, thus uncovering the wave-like nature of electric and magnetic fields, and their
symmetry. Because the speed of EM waves predicted by the wave equation coincided with the
measured speed of light, Maxwell concluded that light itself is an EM wave.

According to Maxwell's equations, a time-varying electric field generates a time-varying


magnetic field and vice versa. Therefore, as an oscillating electric field generates an oscillating
magnetic field, the magnetic field in turn generates an oscillating electric field, and so on. These
oscillating fields together form a propagating electromagnetic wave.
A quantum theory of the interaction between electromagnetic radiation and matter such as
electrons is described by the theory of quantum electrodynamics.
[edit] Properties

Electromagnetic waves can be imagined as a self-propagating transverse oscillating


wave of electric and magnetic fields. This diagram shows a plane linearly polarized
wave propagating from right to left. The electric field is in a vertical plane and the
magnetic field in a horizontal plane.

The physics of electromagnetic radiation is electrodynamics. Electromagnetism is the physical


phenomenon associated with the theory of electrodynamics. Electric and magnetic fields obey
the properties of superposition. Thus, a field due to any particular particle or time-varying
electric or magnetic field contributes to the fields present in the same space due to other causes.
Further, as they are vector fields, all magnetic and electric field vectors add together according to
vector addition. For example, in optics two or more coherent lightwaves may interact and by
constructive or destructive interference yield a resultant irradiance deviating from the sum of the
component irradiances of the individual lightwaves.
Since light is an oscillation it is not affected by travelling through static electric or magnetic
fields in a linear medium such as a vacuum. However in nonlinear media, such as some crystals,
interactions can occur between light and static electric and magnetic fields these interactions
include the Faraday effect and the Kerr effect.
In refraction, a wave crossing from one medium to another of different density alters its speed
and direction upon entering the new medium. The ratio of the refractive indices of the media
determines the degree of refraction, and is summarized by Snell's law. Light disperses into a
visible spectrum as light passes through a prism because of the wavelength dependent refractive
index of the prism material (Dispersion).
EM radiation exhibits both wave properties and particle properties at the same time (see waveparticle duality). Both wave and particle characteristics have been confirmed in a large number
of experiments. Wave characteristics are more apparent when EM radiation is measured over
relatively large timescales and over large distances while particle characteristics are more evident
when measuring small timescales and distances. For example, when electromagnetic radiation is

absorbed by matter, particle-like properties will be more obvious when the average number of
photons in the cube of the relevant wavelength is much smaller than 1. Upon absorption of light,
it is not too difficult to experimentally observe non-uniform deposition of energy. Strictly
speaking, however, this alone is not evidence of "particulate" behavior of light, rather it reflects
the quantum nature of matter.[1]
There are experiments in which the wave and particle natures of electromagnetic waves appear in
the same experiment, such as the self-interference of a single photon. True single-photon
experiments (in a quantum optical sense) can be done today in undergraduate-level labs.[2] When
a single photon is sent through an interferometer, it passes through both paths, interfering with
itself, as waves do, yet is detected by a photomultiplier or other sensitive detector only once.
[edit] Wave model

Electromagnetic radiation is a transverse wave meaning that the oscillations of the waves are
perpendicular to the direction of energy transfer and travel. An important aspect of the nature of
light is frequency. The frequency of a wave is its rate of oscillation and is measured in hertz, the
SI unit of frequency, where one hertz is equal to one oscillation per second. Light usually has a
spectrum of frequencies which sum together to form the resultant wave. Different frequencies
undergo different angles of refraction.
A wave consists of successive troughs and crests, and the distance between two adjacent crests or
troughs is called the wavelength. Waves of the electromagnetic spectrum vary in size, from very
long radio waves the size of buildings to very short gamma rays smaller than atom nuclei.
Frequency is inversely proportional to wavelength, according to the equation:

where v is the speed of the wave (c in a vacuum, or less in other media), f is the frequency and
is the wavelength. As waves cross boundaries between different media, their speeds change but
their frequencies remain constant.
Interference is the superposition of two or more waves resulting in a new wave pattern. If the
fields have components in the same direction, they constructively interfere, while opposite
directions cause destructive interference.
The energy in electromagnetic waves is sometimes called radiant energy.
[edit] Particle model
See also: Quantization (physics) and Quantum optics

Because energy of an EM interaction is quantized, EM waves are emitted and absorbed as


discrete packets of energy, or quanta, called photons.[3] Because photons are emitted and
absorbed by charged particles, they act as transporters of energy, and are associated with waves
with frequency proportional to the energy carried. The energy per photon can be related to the
frequency via the PlanckEinstein equation:[4]

where E is the energy, h is Planck's constant, and f is frequency. The energy is commonly
expressed in the unit of electronvolt (eV). This photon-energy expression is a particular case of
the energy levels of the more general electromagnetic oscillator, whose average energy, which is
used to obtain Planck's radiation law, can be shown to differ sharply from that predicted by the
equipartition principle at low temperature, thereby establishes a failure of equipartition due to
quantum effects at low temperature.[5]
As a photon is absorbed by an atom, it excites the atom, elevating an electron to a higher energy
level. If the energy is great enough, so that the electron jumps to a high enough energy level, it
may escape the positive pull of the nucleus and be liberated from the atom in a process called
photoionisation. Conversely, an electron that descends to a lower energy level in an atom emits a
photon of light equal to the energy difference. Since the energy levels of electrons in atoms are
discrete, each element emits and absorbs its own characteristic frequencies.
Together, these effects explain the emission and absorption spectra of light. The dark bands in the
absorption spectrum are due to the atoms in the intervening medium absorbing different
frequencies of the light. The composition of the medium through which the light travels
determines the nature of the absorption spectrum. For instance, dark bands in the light emitted by
a distant star are due to the atoms in the star's atmosphere. These bands correspond to the
allowed energy levels in the atoms. A similar phenomenon occurs for emission. As the electrons
descend to lower energy levels, a spectrum is emitted that represents the jumps between the
energy levels of the electrons. This is manifested in the emission spectrum of nebulae. Today,
scientists use this phenomenon to observe what elements a certain star is composed of. It is also
used in the determination of the distance of a star, using the red shift.
[edit] Speed of propagation
Main article: Speed of light

Any electric charge which accelerates, or any changing magnetic field, produces electromagnetic
radiation. Electromagnetic information about the charge travels at the speed of light. Accurate
treatment thus incorporates a concept known as retarded time (as opposed to advanced time,
which is not physically possible in light of causality), which adds to the expressions for the
electrodynamic electric field and magnetic field. These extra terms are responsible for
electromagnetic radiation. When any wire (or other conducting object such as an antenna)
conducts alternating current, electromagnetic radiation is propagated at the same frequency as
the electric current. At the quantum level, electromagnetic radiation is produced when the
wavepacket of a charged particle oscillates or otherwise accelerates. Charged particles in a
stationary state do not move, but a superposition of such states may result in oscillation, which is
responsible for the phenomenon of radiative transition between quantum states of a charged
particle.
Depending on the circumstances, electromagnetic radiation may behave as a wave or as particles.
As a wave, it is characterized by a velocity (the speed of light), wavelength, and frequency.
When considered as particles, they are known as photons, and each has an energy related to the

frequency of the wave given by Planck's relation E = h, where E is the energy of the photon, h
= 6.626 1034 Js is Planck's constant, and is the frequency of the wave.
One rule is always obeyed regardless of the circumstances: EM radiation in a vacuum always
travels at the speed of light, relative to the observer, regardless of the observer's velocity. (This
observation led to Albert Einstein's development of the theory of special relativity.)
In a medium (other than vacuum), velocity factor or refractive index are considered, depending
on frequency and application. Both of these are ratios of the speed in a medium to speed in a
vacuum.

Thermal radiation and electromagnetic radiation as a form


of heat
Main article: Thermal radiation

The basic structure of matter involves charged particles bound together in many different ways.
When electromagnetic radiation is incident on matter, it causes the charged particles to oscillate
and gain energy. The ultimate fate of this energy depends on the situation. It could be
immediately re-radiated and appear as scattered, reflected, or transmitted radiation. It may also
get dissipated into other microscopic motions within the matter, coming to thermal equilibrium
and manifesting itself as thermal energy in the material. With a few exceptions such as
fluorescence, harmonic generation, photochemical reactions and the photovoltaic effect,
absorbed electromagnetic radiation simply deposits its energy by heating the material. This
happens both for infrared and non-infrared radiation. Intense radio waves can thermally burn
living tissue and can cook food. In addition to infrared lasers, sufficiently intense visible and
ultraviolet lasers can also easily set paper afire. Ionizing electromagnetic radiation can create
high-speed electrons in a material and break chemical bonds, but after these electrons collide
many times with other atoms in the material eventually most of the energy gets downgraded to
thermal energy, this whole process happening in a tiny fraction of a second. That infrared
radiation is a form of heat and other electromagnetic radiation is not, is a widespread
misconception in physics. Any electromagnetic radiation can heat a material when it is absorbed.
The inverse or time-reversed process of absorption is responsible for thermal radiation. Much of
the thermal energy in matter consists of random motion of charged particles, and this energy can
be radiated away from the matter. The resulting radiation may subsequently be absorbed by
another piece of matter, with the deposited energy heating the material. Radiation is an important
mechanism of heat transfer.
The electromagnetic radiation in an opaque cavity at thermal equilibrium is effectively a form of
thermal energy, having maximum radiation entropy. The thermodynamic potentials of
electromagnetic radiation can be well-defined as for matter. Thermal radiation in a cavity has
energy density (see Planck's Law) of

Differentiating the above with respect to temperature, we may say that the electromagnetic
radiation field has an effective volumetric heat capacity given by

Electromagnetic spectrum
Main article: Electromagnetic spectrum

Electromagnetic spectrum with light highlighted

Legend:
= Gamma rays
HX = Hard X-rays
SX = Soft X-Rays
EUV = Extreme ultraviolet
NUV = Near ultraviolet
Visible light
NIR = Near infrared
MIR = Moderate infrared
FIR = Far infrared
Radio waves:
EHF = Extremely high frequency (Microwaves)
SHF = Super high frequency (Microwaves)
UHF = Ultrahigh frequency
VHF = Very high frequency
HF = High frequency
MF = Medium frequency
LF = Low frequency
VLF = Very low frequency

VF = Voice frequency
ULF = Ultra low frequency
SLF = Super low frequency
ELF = Extremely low frequency

Generally, EM radiation (the designation 'radiation' excludes static electric and magnetic and
near fields) is classified by wavelength into radio, microwave, infrared, the visible region we
perceive as light, ultraviolet, X-rays and gamma rays. Arbitrary electromagnetic waves can
always be expressed by Fourier analysis in terms of sinusoidal monochromatic waves which can
be classified into these regions of the spectrum.
The behavior of EM radiation depends on its wavelength. Higher frequencies have shorter
wavelengths, and lower frequencies have longer wavelengths. When EM radiation interacts with
single atoms and molecules, its behavior depends on the amount of energy per quantum it carries.
Spectroscopy can detect a much wider region of the EM spectrum than the visible range of
400 nm to 700 nm. A common laboratory spectroscope can detect wavelengths from 2 nm to
2500 nm. Detailed information about the physical properties of objects, gases, or even stars can
be obtained from this type of device. It is widely used in astrophysics. For example, hydrogen
atoms emit radio waves of wavelength 21.12 cm.
Soundwaves are not electromagnetic radiation. At the lower end of the electromagnetic
spectrum, about 20 Hz to about 20 kHz, are frequencies that might be considered in the audio
range. However, electromagnetic waves cannot be directly perceived by human ears. Sound
waves are the oscillating compression of molecules. To be heard, electromagnetic radiation must
be converted to air pressure waves, or if the ear is submerged, water pressure waves.
[edit] Light
Main article: Light

EM radiation with a wavelength between approximately 400 nm and 700 nm is directly detected
by the human eye and perceived as visible light. Other wavelengths, especially nearby infrared
(longer than 700 nm) and ultraviolet (shorter than 400 nm) are also sometimes referred to as
light, especially when visibility to humans is not relevant.
If radiation having a frequency in the visible region of the EM spectrum reflects off of an object,
say, a bowl of fruit, and then strikes our eyes, this results in our visual perception of the scene.
Our brain's visual system processes the multitude of reflected frequencies into different shades
and hues, and through this not-entirely-understood psychophysical phenomenon, most people
perceive a bowl of fruit.
At most wavelengths, however, the information carried by electromagnetic radiation is not
directly detected by human senses. Natural sources produce EM radiation across the spectrum,
and our technology can also manipulate a broad range of wavelengths. Optical fiber transmits
light which, although not suitable for direct viewing, can carry data that can be translated into
sound or an image. To be meaningful both transmitter and receiver must use some agreed-upon

encoding system - especially so if the transmission is digital as opposed to the analog nature of
the waves.
[edit] Radio waves
Main article: Radio waves

Radio waves can be made to carry information by varying a combination of the amplitude,
frequency and phase of the wave within a frequency band.
When EM radiation impinges upon a conductor, it couples to the conductor, travels along it, and
induces an electric current on the surface of that conductor by exciting the electrons of the
conducting material. This effect (the skin effect) is used in antennas. EM radiation may also
cause certain molecules to absorb energy and thus to heat up; this is exploited in microwave
ovens. Radio waves are not ionizing radiation, as the energy per photon is too small.

Derivation
This article's tone or style may not be appropriate for Wikipedia.
Specific concerns may be found on the talk page. See Wikipedia's guide to
writing better articles for suggestions. (April 2010)

Electromagnetic waves as a general phenomenon were predicted by the classical laws of


electricity and magnetism, known as Maxwell's equations. Inspection of Maxwell's equations
without sources (charges or currents) results in, along with the possibility of nothing happening,
nontrivial solutions of changing electric and magnetic fields. Beginning with Maxwell's
equations in free space:

where
is a vector differential operator (see Del).

One solution,
,

is trivial.
For a more useful solution, we utilize vector identities, which work for any vector, as follows:

To see how we can use this, take the curl of equation (2):

Evaluating the left hand side:

where we simplified the above by using equation (1).

Evaluate the right hand side:

Equations (6) and (7) are equal, so this results in a vector-valued differential equation for the
electric field, namely

Applying a similar pattern results in similar differential equation for the magnetic field:

These differential equations are equivalent to the wave equation:

where
c0 is the speed of the wave in free space and

f describes a displacement

Or more simply:

where

is d'Alembertian:

Notice that in the case of the electric and magnetic fields, the speed is:

Which, as it turns out, is the speed of light in vacuum. Maxwell's equations have unified the
vacuum permittivity 0, the vacuum permeability 0, and the speed of light itself, c0. Before this
derivation it was not known that there was such a strong relationship between light and
electricity and magnetism.
But these are only two equations and we started with four, so there is still more information
pertaining to these waves hidden within Maxwell's equations. Let's consider a generic vector
wave for the electric field.

Here

is the constant amplitude, f is any second differentiable function,

direction of propagation, and is a position vector. We observe that


generic solution to the wave equation. In other words

is a unit vector in the


is a

for a generic wave traveling in the

direction.

This form will satisfy the wave equation, but will it satisfy all of Maxwell's equations, and with
what corresponding magnetic field?

The first of Maxwell's equations implies that electric field is orthogonal to the direction the wave
propagates.

The second of Maxwell's equations yields the magnetic field. The remaining equations will be
satisfied by this choice of
.
Not only are the electric and magnetic field waves traveling at the speed of light, but they have a
special restricted orientation and proportional magnitudes, E0 = c0B0, which can be seen
immediately from the Poynting vector. The electric field, magnetic field, and direction of wave
propagation are all orthogonal, and the wave propagates in the same direction as
.
From the viewpoint of an electromagnetic wave traveling forward, the electric field might be
oscillating up and down, while the magnetic field oscillates right and left; but this picture can be
rotated with the electric field oscillating right and left and the magnetic field oscillating down
and up. This is a different solution that is traveling in the same direction. This arbitrariness in the
orientation with respect to propagation direction is known as polarization. On a quantum level, it
is described as photon polarization.
More general forms of the second order wave equations given above are available, allowing for
both non-vacuum propagation media and sources. A great many competing derivations exist, all
with varying levels of approximation and intended applications. One very general example is a
form of the electric field equation,[6] which was factorized into a pair of explicitly directional
wave equations, and then efficiently reduced into a single uni-directional wave equation by
means of a simple slow-evolution approximation.

See also

Antenna (radio)
Antenna measurement

Bioelectromagnetism

Bolometer

Control of electromagnetic radiation

Electromagnetic field

Electromagnetic pulse

Electromagnetic radiation and health

Electromagnetic spectrum

Electromagnetic wave equation

Evanescent wave coupling

Finite-difference time-domain method

Helicon

Impedance of free space

Light

Maxwell's equations

Near and far field

Radiant energy

Radiation reaction

Risks and benefits of sun exposure

Sinusoidal plane-wave solutions of the electromagnetic wave equation

References
1. ^ [1]
2. ^ [2]
3. ^ Weinberg, S. (1995). The Quantum Theory of Fields. 1. Cambridge
University Press. pp. 1517. ISBN 0-521-55001-7.
4. ^ Paul M. S. Monk (2004). Physical Chemistry. John Wiley and Sons. p. 435.
ISBN 9780471491804. http://books.google.com/?
id=LupAi35QjhoC&pg=PA435&dq=%22planck+einstein+equation%22.
5. ^ Vu-Quoc, L., Configuration integral (statistical mechanics), 2008.
6. ^ Kinsler, P. (2010). "Optical pulse propagation with minimal approximations"
(reprint). Phys. Rev. A 81: 013819. doi:10.1103/PhysRevA.81.013819].
http://arxiv.org/abs/0810.5689.

Hecht, Eugene (2001). Optics (4th ed.). Pearson Education. ISBN 0-80538566-5.

Serway, Raymond A.; Jewett, John W. (2004). Physics for Scientists and
Engineers (6th ed.). Brooks Cole. ISBN 0-534-40842-7.

Tipler, Paul (2004). Physics for Scientists and Engineers: Electricity,


Magnetism, Light, and Elementary Modern Physics (5th ed.). W. H. Freeman.
ISBN 0-7167-0810-8.

Reitz, John; Milford, Frederick; Christy, Robert (1992). Foundations of


Electromagnetic Theory (4th ed.). Addison Wesley. ISBN 0-201-52624-7.

Jackson, John David (1999). Classical Electrodynamics (3rd ed.). John Wiley &
Sons. ISBN 0-471-30932-X.

Allen Taflove and Susan C. Hagness (2005). Computational Electrodynamics:


The Finite-Difference Time-Domain Method, 3rd ed. Artech House Publishers.
ISBN 1-58053-832-0.

You might also like