Вы находитесь на странице: 1из 5

7/7/14 Lasers | Photonics Handbook

www.photonics.com/EDU/Handbook.aspx?Tag=Lasers+%26+Light+Sources&AID=25161 1/5
Lasers
Understanding the Basics
Coherent, Inc.
Although lasers range from quantum-dot to football-field size and utilize materials from gases to solids, the underlying operating principles are always
the same. This article provides the basic information about how and why lasers work.
When the laser was first demonstrated in 1960, it sparked a wave of public interest. Soon, however, many scientists and engineers dismissed the laser as a
solution without a problem. Time has proved the critics very wrong. From communications to construction, laser technology has become a part of everyday
life.
All light sources convert input energy into light. In the case of the laser, the input, or pump, energy can take many forms, the two most common being
optical and electrical. For optical pumping, the energy source may even be another laser.
In a conventional (incoherent) light source, each atom excited by input energy randomly emits a single photon according to a given statistical probability.
This produces radiation in all directions with a spread of wavelengths and no interrelationships among individual photons. This is called spontaneous
emission.
Figure 1. Spontaneous emission is a random process, whereas stimulated emission produces photons with
identical properties.
Einstein predicted that excited atoms also could convert their stored energy into light by a process called
stimulated emission. Here, an excited atom first produces a photon by spontaneous emission. When this
photon reaches another excited atom, the interaction prompts that atom to emit a second photon (Figure 1).
This process has two important characteristics. First, it is multiplicative one photon becomes two. If these
two photons interact with two other excited atoms, this will yield a total of four photons, and so forth. Most
important, these two photons have identical properties: wavelength, direction, phase and polarization. This
ability to amplify light is termed optical gain, and a wide range of solid, liquid and gas phase materials have
been discovered that exhibit gain.
The laser cavity
The laser cavity, or resonator, is at the heart of the system. The cavity defines a unique axis with very high
optical gain, which becomes the beam direction. This axis usually is defined in two ways. First, the lasers shape ensures that the gain medium is longer
along one axis, often as a long thin cylinder, as is typical for gas lasers. A more extreme example of a uniquely long gain axis is the fiber laser. This is
sufficient for some high-gain devices such as excimers. But for most, it is necessary to further enhance the gain along this axis using cavity mirrors that
produce feedback (Figure 2).
Figure 2. In the prototypical gas laser, the gain medium has a long, thin
cylindrical shape. The cavity is defined by two mirrors. One is partially reflecting
and allows the output beam to escape.
The simplest cavity is defined by two mirrors a total reflector and a partial
reflector whose reflectance can vary between 50 and 99 percent. Light bounces
back and forth between these mirrors, gaining intensity with each pass through
the gain medium. Because some of this light escapes the cavity, or oscillator,
through the partial reflector (output coupler), a stable equilibrium condition is
reached quickly. The output beam is the light that escapes through the output coupler.
In the ideal laser, all the photons in the output beam are identical. This imparts several unique properties: directionality, monochromaticity, coherence and
brightness.
Monochromaticity A photons energy determines its wavelength through the relationship E = hc/, where c is the speed of light, h is Plancks constant and
is wavelength. If our ideal laser emits all photons with the same energy, and thus the same wavelength, it is said to be monochromatic. Many applications
are dependent on monochromaticity. For example, in telecommunications, several lasers at different wavelengths transmit multiple streams of data down the
same fiber without crosstalk.
Coherence Besides being the same wavelength, the photons that make up a laser beam are all in phase (Figure 3). In the ideal case, the laser acts as one
long, continuous, intense lightwave. This enables a host of applications that rely on optical interference. For example, the surface of precision lenses and
mirrors is measured using laser interferometers. The coherent light beam acts as an ultrafine ruler, where the wavelength of light fulfills the dimensional
role.
Figure 3. Laser light differs from conventional light in that all the lightwaves are in phase with each other.
Directionality and brightness The most obvious visible difference between lasers and conventional light sources is that laser light travels in the same
direction as an intense beam. Brightness is defined as the amount of light leaving the source per unit of surface area. Because a lasers photons have
identical vector properties, they act as if they are coming from the same point in space. The ideal laser thus acts as a true point source with extremely high
brightness.
Related Searches
Optical Fiber
Optical System
Atomic Force
Microscope
Light-Emitting Diodes
Fiber Optic
Communication
Electron Microscopy
Electron Microscope
Transmission
Electron Microscope
?
BetterBrowse
7/7/14 Lasers | Photonics Handbook
www.photonics.com/EDU/Handbook.aspx?Tag=Lasers+%26+Light+Sources&AID=25161 2/5
This combination of directionality and brightness has two consequences. The beam can be projected over great
distances, and it can be focused to a very small spot. In the ideal case, the divergence of a collimated beam or the
size of the focused spot are limited only by diffraction an inescapable property of light. This is referred to as a
diffraction-limited beam.
Continuous-wave lasers
Lasers can be divided into three main categories continuous wave (CW), pulsed and ultrafast.
As their name suggests, continuous wave lasers produce a continuous, uninterrupted output. The exact
wavelength(s) at which this occurs is primarily determined by three factors: the gain bandwidth of the lasing
medium, the spectral characteristics of the cavity optics and the longitudinal modes of the resonator.
Many laser materials in fact have several wavelengths (or laser lines) at which emission occurs. Also, several
factors (such as the Doppler effect in the moving atoms of a gas ) typically broaden the wavelength bandwidth of
the gain at each of these various lines.
The first step in determining at which wavelength the laser will operate is to use cavity mirrors that are highly reflective only at the desired wavelength(s).
This suppresses lasing at other lines. However, even a single laser line actually covers a band of wavelengths.
Figure 4. A resonant cavity supports only modes that meet the resonance condition, N = 2/ (cavity
length). The output of a CW laser is defined by the overlap of the gain bandwidth and these resonant
cavity modes.
The specific wavelengths of output within this gain bandwidth are determined by the longitudinal
modes of the cavity. Figure 4 shows the basic principles of the resonant two-mirror cavity, the most
basic design. To sustain gain as light travels back and forth between the mirrors, the waves must
remain in phase, which means that the cavity round-trip distance must be an exact multiple of the
wavelength.
N = 2/(cavity length)
where is the laser wavelength and N is an integer called the mode number it is usually a very large
integer, since the wavelength of light is so much smaller than a typical cavity length. In a helium-neon
laser, for example, the red output wavelength is 0.633 m, yet the typical cavity length is 15 to 50
cm. Wavelengths that satisfy this resonance equation are called longitudinal cavity modes. The actual
output wavelengths are at the cavity modes that fall within the gain bandwidth, as shown in Figure 4.
This is called multi-longitudinal mode operation.
The cavity also controls the transverse modes, or intensity cross sections. The ideal beam has a
symmetric cross section: The intensity is greater in the middle and tails off at the edges. This is
called the TEM
00
output mode. Lasers can produce many other TEM modes, a few of which are shown
in Figure 5. Typically, laser output is specified as what percentage of the total beam intensity is in the
form of the TEM
00
.
Figure 5. Lasers can emit any number of transverse
modes, of which the TEM
00
usually is most desirable.
A laser that produces multiple longitudinal modes has limited coherence different wavelengths cannot
stay in phase over extended distances. Applications such as holography, which demand excellent
coherence, often require a single longitudinal mode laser. For some laser types, single-mode output is
achieved with a very short resonant cavity; this makes the mode spacing larger than the gain bandwidth
and only one mode lases. Generally, though, a filtering element that preferentially passes only one mode is
inserted into the cavity. The most common type of filter is called an etalon.
Various liquid and solid-state lasers have broad bandwidths that cover tens of nanometers. Examples
include dye and Ti:sapphire lasers. Rather than being a disadvantage, this has allowed the development of
tunable and ultrafast lasers. Creating a tunable CW laser involves including an extra filtering element in
the cavity usually a birefringent (or Lyot) filter. The birefringent filter does two things: It narrows the
bandwidth and, by rotating the filter, allows smooth bandwidth tuning, or in the case of optically pumped
semiconductor lasers (OPSLs), it allows the final output wavelength to be exactly set to match the end
users specifications.
Although this sounds complicated, laser operation is remarkably simple. High-end CW lasers include an on-
board computer or microprocessor. This automatically controls the additional cavity elements while
simultaneously maintaining optimum alignment of the mirrors.
Pulsed lasers
Some materials ruby, rare-gas halogen excimers, such as ArF and XeCl sustain laser action for only a
brief period and form the basis of pulsed lasers. If the pulse duration is sufficiently long (microseconds),
the laser can be designed much like a CW laser. However, many pulsed lasers are designed for short pulse
duration; e.g., a few nanoseconds (10
9
s). In each pulse, the light has time for very few round-trips in the
cavity. The resonant cavity designs described so far cannot control such a laser: The pulse dies before
equilibrium conditions are reached.
While two mirrors are still used in pulsed lasers for defining the direction of highest gain, they do not act
7/7/14 Lasers | Photonics Handbook
www.photonics.com/EDU/Handbook.aspx?Tag=Lasers+%26+Light+Sources&AID=25161 3/5
as a resonant cavity. Instead, the usual method of controlling and tuning wavelength is a diffraction grating (Figure 6). Some pulsed lasers, such as Nd:YAG
(neodymium yttrium aluminum garnet) can be operated with a Q-switch, an intracavity device that acts as a fast optical gate. Light cannot pass it unless it is
activated, usually by a high-voltage pulse. Initially, the switch is closed and energy is allowed to build up in the laser material. Then at the optimum time, the
switch is opened and the stored energy is released as a very short pulse. This can shorten the normal pulse duration by several orders of magnitude. The peak
power of a pulsed laser is proportional to pulse energy/pulse duration. Q-switching, therefore, has an added benefit of increasing peak power by several
orders of magnitude. This effect enables neodymium (Nd)-based solid state lasers of only modest average power to machine tougher materials such as
glasses and metals.
Figure 6. The wavelength of a pulsed laser usually is controlled with a diffraction grating. Rotating
the angle of this device tunes the wavelength.
The wavelength purity of Q-switched lasers can be difficult to control because of the combination of
the high peak power and the short pulse duration. However, this can be solved by stacking two or
more lasers in series: a low-power, well-controlled oscillator followed by one or more amplifiers. For
even higher performance, the oscillator itself is sometimes seeded by another low-power laser, such as a wavelength-stabilized laser diode.
A much less common pulsing mechanism is called cavity-dumping and is used where the laser material cannot store enough gain for stable Q-switched
operation. A cavity dumped laser has end mirrors with nominal 100-percent efficiency in order to maximize the circulating intra-cavity power. An intracavity
switch, either an acousto-optic deflector or tillable mirror is then flipped to allow all the trapped energy to depart the cavity in a single round-trip time
interval. Pulse energies achievable with cavity dumping are much lower than Q-switching where the power is stored as gain in the laser medium.
Ultrafast lasers
CW lasers can produce many longitudinal modes, and if the cavity is pulsed or modulated, it is possible to lock the phase of these modes together. The
resultant interference causes the traveling lightwaves inside the cavity to collapse into a very short pulse. Every time this pulse reaches the output coupler,
the laser emits a part of this pulse. The pulse repetition rate is determined by the time it takes the pulse to make one trip around the cavity.
It turns out that the more modes that interfere, the shorter the pulse duration. In other words, the pulse duration is inversely proportional to the bandwidth
of the laser gain material. This explains why the materials used for broadly tunable lasers produce the shortest mode-locked pulses. The most popular
ultrafast laser material is titanium-doped sapphire or Ti:sapphire; turnkey commercial Ti:sapphire lasers now routinely deliver pulses as short as 10 fs (20 x
10
15
s), with typical repetition rates around 100 MHz. and peak powers approaching 1 MW. This can be amplified to the terawatt level in custom commercial
products.
With the CW power condensed into a mode-locked pulse, the result is high peak power even for modest devices. Furthermore, a regenerative amplifier can
boost the peak power of an ultrafast laser by orders of magnitude. These lasers can be optimized for high repetition rates (300 kHz) or high peak power
(20 x 10
12
W) the highest peak power delivered by any class of commercial laser.
Specialized ultrafast lasers CEP stabilization
In a mode-locked laser, interference between the myriad longitudinal modes means that in the time domain, the output collapses from a continuous-wave
light source to a series of short pulses separated by the time it takes for light to travel around the cavity. But in the frequency (1/wavelength) regime, the
pulse still consists of these myriad individual modes. An intensity vs. frequency plot of these modes looks like a comb (Figure 4) and are often referred to as
a frequency comb. In recent years there has been fast growing interest, and even a Nobel prize, for using this type of comb as a tool for high resolution
spectroscopy with unprecedented absolute precision. When using a laser in this way, the modes act as ultrahigh precision frequency (wavelength) fiducials.
This requires a specialized laser operating approach called carrier envelope phase (CEP) stabilization.
In simple terms, each mode has a frequency given by
= c/ = Nc/(2x cavity length)
Where c is the speed of light and N is the mode number. But in real lasers, this is not exactly true. Thats because the speed of light in any material other
than a vacuum is actually defined by two separate velocity components called group velocity and phase velocity. The group velocity is the effective speed with
which the light travels and the phase velocity is the speed of the actual wavefronts the waves that are electric-field oscillations. The difference in group and
phase velocity is a direct function of the refractive index of the material(s) the light is passing through.
In a mode-locked laser this can be seen because it has the effect of adding an offset to the above formula
= Nc/(2x cavity length) + offset
This is called the carrier offset frequency or CEO. In a spectroscopy measurement, the full utility of using the comb of modes as reference frequencies means
that this offset must be held constant throughout the spectral acquisition time. But even extremely minor fluctuations and drifts in the optical properties of
the laser cavity (e.g., noise in the laser pumping the Ti:sapphire crystal) can cause significant shifts in the CEO value. This can be successfully addressed by
using a feedback loop that measures the CEO value and then adjusts one of the cavity optics to hold the CEO value constant. The most robust CEO
measurement tool for this purpose is the so-called 1f 2f interferometer whose operation has been well described elsewhere.
If the CEO is held at zero, the modes are exact multiples of the inverse of the cavity length, enabling absolute measurement of high resolution (hyperfine)
spectrum parameters. But there is another advantage to holding the CEO value at zero as can be seen in Figure 7. Specifically, in the time domain, when the
CEO is zero, the peak of the overall electric field oscillation coincides with the amplitude envelope of the ultrafast pulse. Clearly this results in the highest
possible electric field peak value. This is a critical advantage for applications involving multiple nonlinear processes, i.e., optical processes whose
efficiencies have a high order dependence on the peak electric field of the laser. At this time, the most spectacular of these applications are in attosecond
physics using ultrashort x-ray pulses. These short pulses offer chemists and physicists the first tool capable of probing electron motion, in contrast to slower
spectroscopy methods that can only follow the motion of the much slower nuclei.
Figure 7. In typical ultrafast laser operation, the electric field oscillation does not have a fixed phase relationship with the pulse envelope, as shown here.
The goal of CEP stabilization is to fix the phase relationship between the overall pulse envelope and the underlying electric field oscillation.
CEP stabilization requires a very low noise system, including a low-noise CW laser to pump the Ti:sapphire oscillator. In addition, the oscillator and amplifier
7/7/14 Lasers | Photonics Handbook
www.photonics.com/EDU/Handbook.aspx?Tag=Lasers+%26+Light+Sources&AID=25161 4/5
must both incorporate feedback mechanisms that adjust cavity dispersion in real time. Fortunately, turnkey
commercial laser systems are now available with all these attributes, enabling scientists to focus on their
experiments rather than on the nuances of CEP stabilization.
Frequency doubling
Even with the wealth of commercially available lasers, it is not always possible to find one that exactly matches
an application. Fortunately, the required wavelength often can be generated using frequency doubling, or
shifting, or with an optical parametric oscillator. All these processes are related and are called nonlinear
phenomena since they depend nonlinearly on the lasers peak power.
In simple terms, when an intense and/or tightly focused laser beam passes through a suitable condensed phase
such as a liquid, solid crystal or even dense gas jet, its oscillating electric field may interact with the electrons of
the atoms or molecules in several ways. One of these mechanisms serves to distort the electron cloud thereby
polarizing the atoms, i.e., the traveling sine wave creates a temporary refractive index profile that is also a
traveling sine wave. When a photon interacts with this moving refractive index ripple it can gain energy from the
interaction or lose energy from the interaction. This is the basis of frequency-doubling, where the laser frequency
is doubled and the wavelength is thus halved, as well as frequency mixing, where two laser wavelengths combine
to form photons with the sum of their original energies. It can also be used to create long wavelength photons
whose energy is the difference between two different input laser wavelengths. Because these nonlinear
interactions depend on the second or third power of the laser intensity, they work well with pulsed lasers that
have high peak power. They can also be used with CW lasers if the beam is focused and if the nonlinear crystal is
inside the laser cavity (intracavity doubling).
Under most conditions, the light at the new frequency (wavelength) would be destroyed by destructive interference. Thats because it is usually created in a
long (millimeters) crystal using the full length of that crystal. But the phase velocity of the original and new wavelengths is different. So wavelength-shifted
light created at one spot in the crystal is out of phase with that created at another position along the crystal, and so on. This difficulty is overcome by
choosing a crystal temperature and orientation that creates a so-called phase-matching condition where the phase velocity of the fundamental and shifted
light is the same. Details of phase matching are beyond the scope of this review article. But the most common phase-matching mechanisms depend on
birefringence where the phase velocity of light in a crystal depends on the lights polarization angle.
Optical parametric oscillators
The optical parametric oscillator (OPO) represents an area of rapid development in terms of both products and applications thanks to its ability to produce
tunable output anywhere from the mid-UV to the mid-IR. For example, it is proving a powerful tool in the near-infrared for use in deep-tissue microscopy.
An OPO uses a similar nonlinear mechanism in a laserlike cavity to generate two shorter frequencies (i.e., longer wavelengths) from one input frequency. The
shorter of these new wavelengths is called the signal wavelength, and the longer, the idler. The exact values can be smoothly varied by tuning the cavity and
rotating the OPO crystal under microprocessor control. Because of the high power necessary for OPO operation, these devices typically have been limited to
pulsed and ultrafast systems.
Recently, the use of fan-poled nonlinear crystals have enabled compact OPOs that can operate equally well over a wide range of input wavelengths. By using a
tunable pump laser such as a Ti:sapphire laser, this type of OPO can thus be utilized to provide the user with two independently tunable wavelengths. This
can be very useful in several types of microscopy including coherent anti-Stokes Raman spectroscopy (CARS) imaging.
A related device is the optical parametric amplifier, which was developed to provide tunable pulses with much higher pulse energy, e.g., for gas phase
experiments. They are pumped by an ultrafast amplifier usually a regenerative amplifier in order to provide the requisite beam quality.
Common laser types
For many years, the most common CW laser was the helium neon laser, or HeNe. These low-power lasers (a few milliwatts) use an electric discharge to create
a low-pressure plasma in a glass tube; nearly all emit in the red at 633 nm. In recent years, the majority of HeNe applications have switched to visible laser
diodes. Typical applications include bar-code readers, alignment tasks in the construction and the lumber industries, and a host of sighting and pointing
applications from medical surgery to high-energy physics.
In fact, the laser diode has become by far the most common laser type, with truly massive use throughout telecommunications and data storage (e.g., DVDs,
CDs). In a laser diode, current flow creates charge carriers (electrons and holes) in a p-n junction. These combine and emit light through stimulated
emission. Laser diodes are available as single emitters with powers up to tens of watts, and as monolithic linear bars, with numerous individual emitters.
These bars can be assembled into 2-D arrays with total output powers in the kilowatts range. They are used in both CW and pulsed operation for so-called
direct diode applications. But even more importantly, laser diodes now underpin many other types of lasers, where they are used as optical pumps that
perform the initial electrical-to-optical power conversion.
For example, higher power visible CW applications were originally supported by argon-ion and krypton-ion lasers. Based on a plasma discharge tube operating
at high current, these gas-phase lasers are large and very inefficient, generating a large amount of heat which must be actively dissipated. The tube also has
a finite lifetime and thus represents a costly consumable. In most former applications the ion laser was displaced by diode-pumped solid- state (DPSS) lasers.
Here, the gain medium is a neodymium-doped crystal (usually Nd:YAG or Nd:YVO
4
) pumped by one or more laser diodes. The near-IR fundamental at 1064 nm
is then converted to green 532 nm output by the use of an intracavity doubling crystal.
The DPSS laser in turn, has now been challenged by several newer technologies. The most successful of these is the OPSL. Here the gain medium is a large-
area semiconductor laser that is pumped by one or more laser diodes. The OPSL offers numerous advantages, most notably wavelength and power scalablity.
Specifically, these lasers can be designed to operate at virtually any visible wavelength, at last freeing applications from the restrictions of limited legacy
wavelength choices (i.e., 488, 514 and 532 nm). Indeed OPSLs represent a paradigm shift in lasers because they can be designed for the needs of the
application instead of vice versa.
OPSL is now a dominant technology in low-power bioinstrumentation applications, most notably at 488 nm. And the power scalability and inherent low noise of
OPSL technology is now seeing multiwatt green and yellow OPSLs moving strongly into other applications including scientific research, forensics,
ophthalmology, and light shows.
At longer wavelengths, carbon dioxide lasers, which use plasma discharge technology, emit in the mid-infrared around 10 m. Most are CW or pseudo-CW,
with commercial output powers from a few watts to several kilowatts. Another important technology is the fiber laser, which can be operated in CW, Q-
switched and mode-locked formats. Here, laser diodes optically pump a rare-earth doped fiber, which typically emits at about 1 m.
7/7/14 Lasers | Photonics Handbook
www.photonics.com/EDU/Handbook.aspx?Tag=Lasers+%26+Light+Sources&AID=25161 5/5
Nd:YAG, CO
2
, fiber and direct diode lasers are the workhorses of industrial laser applications. Direct diode lasers predominately service low brightness
applications, such as heat treating, cladding and some welding applications. This is because direct diode lasers offer the lowest capital cost of any industrial
laser type, as well as the lowest operating costs, due to their high electrical efficiency.
The advent of slab-discharge technology has allowed the size/power ratio of CO
2
lasers to be greatly scaled down, increasing their utility in subkilowatt
applications. Low-cost waveguide designs also support a healthy market for CO
2
lasers with powers in the tens of watts, primarily in marking and engraving
applications.
Sealed CO
2
lasers and fiber lasers have come to dominate the cutting of metals in the 2- to 4-mm thickness range. Sealed CO
2
is usually the first choice when
both metals and nonmetals must both be processed, while fiber lasers have proved quite successful in certain markets that can benefit from their
combination of high repetition rate, low pulse energy and high brightness. They also excel at metal cutting and welding in the 4- to 6-mm thickness range, as
well as some marking applications. Flowing gas CO
2
lasers still dominate the market for thick metal (> 6 mm) cutting.
Nd:YAG can deliver the high peak power for materials processing applications such as metal welding; in these heavy industrial applications, raw power is
more important than beam quality and for many years these lasers were lamp-pumped. But the ever increasing power and lifetime characteristics of laser
diodes are causing these lasers to switch to diode pumping, i.e., DPSS lasers.
Conversely, lower power Q-switched DPSS lasers are often based on Nd:YVO
4
. These are usually optimized for high beam quality in order for use in
micromachining and microstructuring applications with high repetition rates (up to 250 kHz) to support high throughput processes. They are available with
powers up to tens of watts with a choice of near-infrared (1064 nm), green (532 nm) or UV (355 nm) output. The UV is popular for producing small features in
delicate materials because it can be focused to a very small spot and minimizes peripheral thermal damage. Deep-UV (266 nm) versions are starting to be
used in some applications. But their relatively high cost and the need for specialty beam delivery optics causes many potential applications to rely instead on
355 nm lasers optimized for short pulse duration, which can produce similar results in many materials.
Excimers represent another important pulsed laser technology. They can produce several discrete wavelengths throughout the UV; depending on the gas
combination, emission ranges from 157 to 348 nm. The deep-UV line at 193 nm is the most widely used source for lithography processes in the semiconductor
industry. The 308 nm wavelength is used for annealing silicon in high performance displays. The same wavelength is also key to generating a unique long-
wear surface on the cylinder liners of high performance diesel engines. And finally, excimers have a unique ability to produce high pulse energies up to 1
joule per pulse. This enables direct writing of low-cost electronic circuits for applications such as medical disposables.
Ultrafast lasers for scientific applications are dominated by Ti:sapphire as already described. Ultrafast is also a fast growing technology for micromachining
applications. These lasers are based on fibers, free-space optics or some combination of the two.
And finally, there are many other types of niche and exotic lasers that are beyond the coverage of this overview article. Examples include Raman lasers used
in telecommunications, quantum cascade lasers used in some gas sensing applications, and chemical lasers which tend to be limited to military programs.
Our Magazines
Photonics Spectra
BioPhotonics
EuroPhotonics
Industrial Photonics
Subscribe
Mobile Apps
Products
Latest Products
Photonics Showcase
Photonics Buyers' Guide
Prism Awards for Photonics Innovation
Latest News
Latest News
Editors' Picks
Blogs
Light Matters Weekly Newscast
E-newsletters
RSS Feed
Browse by Subject
Lasers
Optics
Imaging
Microscopy
Spectroscopy
Fiber Optics
Sensors & Detectors
Test & Measurement
Materials, Chemicals & Coatings
Light Sources
Browse by Application
Biophotonics
Industrial
Consumer
Nano
Energy
Defense & Security
Communications
Basic Science
Educational/Reference
EDU.Photonics.com
Photonics Dictionary Plus
Photonics Handbook
Video Gallery
Educational Institutions
Webinars
White Papers
Calendar of Courses & Events
Laser Wavelength Chart
Laser Timeline
Community
Light Exchange Forum
Facebook
Twitter
Site Info and Support
About Us
Contact Us
Advertise / Media Kit
Advertising Contacts
Buyers' Guide Listing
Purchase Buyers' Guide
Purchase Wall Charts
Login
Register
Privacy Policy
Term and Conditions
2014
EDU.Photonics.com is part of Photonics.com and is published by Photonics Media. For more coverage of photonics, visit these pages on Photonics.com.

Вам также может понравиться