Dye laser

Last updated
Close-up of a table-top CW dye laser based on rhodamine 6G, emitting at 580 nm (yellow). The emitted laser beam is visible as faint yellow lines between the yellow window (center) and the yellow optics (upper-right), where it reflects down across the image to an unseen mirror, and back into the dye jet from the lower left corner. The orange dye-solution enters the laser from the left and exits to the right, still glowing from triplet phosphorescence, and is pumped by a 514 nm (blue-green) beam from an argon laser. The pump laser can be seen entering the dye jet, beneath the yellow window. Coherent 899 dye laser.jpg
Close-up of a table-top CW dye laser based on rhodamine 6G, emitting at 580 nm (yellow). The emitted laser beam is visible as faint yellow lines between the yellow window (center) and the yellow optics (upper-right), where it reflects down across the image to an unseen mirror, and back into the dye jet from the lower left corner. The orange dye-solution enters the laser from the left and exits to the right, still glowing from triplet phosphorescence, and is pumped by a 514 nm (blue-green) beam from an argon laser. The pump laser can be seen entering the dye jet, beneath the yellow window.

A dye laser is a laser that uses an organic dye as the lasing medium, usually as a liquid solution. Compared to gases and most solid state lasing media, a dye can usually be used for a much wider range of wavelengths, often spanning 50 to 100 nanometers or more. The wide bandwidth makes them particularly suitable for tunable lasers and pulsed lasers. The dye rhodamine 6G, for example, can be tuned from 635 nm (orangish-red) to 560 nm (greenish-yellow), and produce pulses as short as 16 femtoseconds. [1] Moreover, the dye can be replaced by another type in order to generate an even broader range of wavelengths with the same laser, from the near-infrared to the near-ultraviolet, although this usually requires replacing other optical components in the laser as well, such as dielectric mirrors or pump lasers.

Contents

Dye lasers were independently discovered by P. P. Sorokin and F. P. Schäfer (and colleagues) in 1966. [2] [3]

In addition to the usual liquid state, dye lasers are also available as solid state dye lasers (SSDL). These SSDL lasers use dye-doped organic matrices as gain medium.

Construction

The internal cavity of a linear dye-laser, showing the beam path. The pump laser (green) enters the dye cell from the left. The emitted beam exits to the right (lower yellow beam) through a cavity dumper (not shown). A diffraction grating is used as the high-reflector (upper yellow beam, left side). The two meter beam is redirected several times by mirrors and prisms, which reduce the overall length, expand or focus the beam for various parts of the cavity, and eliminate one of two counter-propagating waves produced by the dye cell. The laser is capable of continuous wave operation or ultrashort picosecond pulses (trillionth of a second, equating to a beam less than
.mw-parser-output .sfrac{white-space:nowrap}.mw-parser-output .sfrac.tion,.mw-parser-output .sfrac .tion{display:inline-block;vertical-align:-0.5em;font-size:85%;text-align:center}.mw-parser-output .sfrac .num,.mw-parser-output .sfrac .den{display:block;line-height:1em;margin:0 0.1em}.mw-parser-output .sfrac .den{border-top:1px solid}.mw-parser-output .sr-only{border:0;clip:rect(0,0,0,0);height:1px;margin:-1px;overflow:hidden;padding:0;position:absolute;width:1px}
1/3 of a millimeter in length). Dye laser alignment intra-cavity beam @ 589nm.jpg
The internal cavity of a linear dye-laser, showing the beam path. The pump laser (green) enters the dye cell from the left. The emitted beam exits to the right (lower yellow beam) through a cavity dumper (not shown). A diffraction grating is used as the high-reflector (upper yellow beam, left side). The two meter beam is redirected several times by mirrors and prisms, which reduce the overall length, expand or focus the beam for various parts of the cavity, and eliminate one of two counter-propagating waves produced by the dye cell. The laser is capable of continuous wave operation or ultrashort picosecond pulses (trillionth of a second, equating to a beam less than 1/3 of a millimeter in length).
A ring dye laser. P-pump laser beam; G-gain dye jet; A-saturable absorber dye jet; M0, M1, M2-planar mirrors; OC-output coupler; CM1 to CM4-curved mirrors. Ring dye2.jpg
A ring dye laser. P-pump laser beam; G-gain dye jet; A-saturable absorber dye jet; M0, M1, M2-planar mirrors; OC–output coupler; CM1 to CM4-curved mirrors.

A dye laser uses a gain medium consisting of an organic dye, which is a carbon-based, soluble stain that is often fluorescent, such as the dye in a highlighter pen. The dye is mixed with a compatible solvent, allowing the molecules to diffuse evenly throughout the liquid. The dye solution may be circulated through a dye cell, or streamed through open air using a dye jet. A high energy source of light is needed to 'pump' the liquid beyond its lasing threshold. A fast discharge flashtube or an external laser is usually used for this purpose. Mirrors are also needed to oscillate the light produced by the dye's fluorescence, which is amplified with each pass through the liquid. The output mirror is normally around 80% reflective, while all other mirrors are usually more than 99.9% reflective. The dye solution is usually circulated at high speeds, to help avoid triplet absorption and to decrease degradation of the dye. A prism or diffraction grating is usually mounted in the beam path, to allow tuning of the beam.

Because the liquid medium of a dye laser can fit any shape, there are a multitude of different configurations that can be used. A Fabry–Pérot laser cavity is usually used for flashtube pumped lasers, which consists of two mirrors, which may be flat or curved, mounted parallel to each other with the laser medium in between. The dye cell is often a thin tube approximately equal in length to the flashtube, with both windows and an inlet/outlet for the liquid on each end. The dye cell is usually side-pumped, with one or more flashtubes running parallel to the dye cell in a reflector cavity. The reflector cavity is often water cooled, to prevent thermal shock in the dye caused by the large amounts of near-infrared radiation which the flashtube produces. Axial pumped lasers have a hollow, annular-shaped flashtube that surrounds the dye cell, which has lower inductance for a shorter flash, and improved transfer efficiency. Coaxial pumped lasers have an annular dye cell that surrounds the flashtube, for even better transfer efficiency, but have a lower gain due to diffraction losses. Flash pumped lasers can be used only for pulsed output applications. [4] [5] [6]

A ring laser design is often chosen for continuous operation, although a Fabry–Pérot design is sometimes used. In a ring laser, the mirrors of the laser are positioned to allow the beam to travel in a circular path. The dye cell, or cuvette, is usually very small. Sometimes a dye jet is used to help avoid reflection losses. The dye is usually pumped with an external laser, such as a nitrogen, excimer, or frequency doubled Nd:YAG laser. The liquid is circulated at very high speeds, to prevent triplet absorption from cutting off the beam. [7] Unlike Fabry–Pérot cavities, a ring laser does not generate standing waves which cause spatial hole burning, a phenomenon where energy becomes trapped in unused portions of the medium between the crests of the wave. This leads to a better gain from the lasing medium. [8] [9]

Operation

The dyes used in these lasers contain rather large organic molecules which fluoresce. Most dyes have a very short time between the absorption and emission of light, referred to as the fluorescence lifetime, which is often on the order of a few nanoseconds. (In comparison, most solid-state lasers have a fluorescence lifetime ranging from hundreds of microseconds to a few milliseconds.) Under standard laser-pumping conditions, the molecules emit their energy before a population inversion can properly build up, so dyes require rather specialized means of pumping. Liquid dyes have an extremely high lasing threshold. In addition, the large molecules are subject to complex excited state transitions during which the spin can be "flipped", quickly changing from the useful, fast-emitting "singlet" state to the slower "triplet" state. [10]

The incoming light excites the dye molecules into the state of being ready to emit stimulated radiation; the singlet state. In this state, the molecules emit light via fluorescence, and the dye is transparent to the lasing wavelength. Within a microsecond or less, the molecules will change to their triplet state. In the triplet state, light is emitted via phosphorescence, and the molecules absorb the lasing wavelength, making the dye partially opaque. Flashlamp-pumped lasers need a flash with an extremely short duration, to deliver the large amounts of energy necessary to bring the dye past threshold before triplet absorption overcomes singlet emission. Dye lasers with an external pump-laser can direct enough energy of the proper wavelength into the dye with a relatively small amount of input energy, but the dye must be circulated at high speeds to keep the triplet molecules out of the beam path. Due to their high absorption, the pumping energy may often be concentrated into a rather small volume of liquid. [11]

Since organic dyes tend to decompose under the influence of light, the dye solution is normally circulated from a large reservoir. [12] The dye solution can be flowing through a cuvette, i.e., a glass container, or be as a dye jet, i.e., as a sheet-like stream in open air from a specially-shaped nozzle. With a dye jet, one avoids reflection losses from the glass surfaces and contamination of the walls of the cuvette. These advantages come at the cost of a more-complicated alignment.

Liquid dyes have very high gain as laser media. The beam needs to make only a few passes through the liquid to reach full design power, and hence, the high transmittance of the output coupler. The high gain also leads to high losses, because reflections from the dye-cell walls or flashlamp reflector cause parasitic oscillations, dramatically reducing the amount of energy available to the beam. Pump cavities are often coated, anodized, or otherwise made of a material that will not reflect at the lasing wavelength while reflecting at the pump wavelength. [11]

A benefit of organic dyes is their high fluorescence efficiency. The greatest losses in many lasers and other fluorescence devices is not from the transfer efficiency (absorbed versus reflected/transmitted energy) or quantum yield (emitted number of photons per absorbed number), but from the losses when high-energy photons are absorbed and reemitted as photons of longer wavelengths. Because the energy of a photon is determined by its wavelength, the emitted photons will be of lower energy; a phenomenon called the Stokes shift. The absorption centers of many dyes are very close to the emission centers. Sometimes the two are close enough that the absorption profile slightly overlaps the emission profile. As a result, most dyes exhibit very small Stokes shifts and consequently allow for lower energy losses than many other laser types due to this phenomenon. The wide absorption profiles make them particularly suited to broadband pumping, such as from a flashtube. It also allows a wide range of pump lasers to be used for any certain dye and, conversely, many different dyes can be used with a single pump laser. [10]

CW dye lasers

Continuous-wave (CW) dye lasers [13] often use a dye jet. CW dye-lasers can have a linear or a ring cavity, and provided the foundation for the development of femtosecond lasers.

Narrow linewidth dye lasers

Multiple prisms expand the beam in one direction, providing better illumination of a diffraction grating. Depending on the angle unwanted wavelengths are dispersed, so are used to tune the output of a dye laser, often to a linewidth of a fraction of an angstrom. Dual prism expander and grating reflector for a dye laser.jpg
Multiple prisms expand the beam in one direction, providing better illumination of a diffraction grating. Depending on the angle unwanted wavelengths are dispersed, so are used to tune the output of a dye laser, often to a linewidth of a fraction of an angstrom.

Dye lasers' emission is inherently broad. However, tunable narrow linewidth emission has been central to the success of the dye laser. In order to produce narrow bandwidth tuning these lasers use many types of cavities and resonators which include gratings, prisms, multiple-prism grating arrangements, and etalons. [14]

The first narrow linewidth dye laser, introduced by Hänsch, used a Galilean telescope as beam expander to illuminate the diffraction grating. [15] Next were the grazing-incidence grating designs [16] [17] and the multiple-prism grating configurations. [18] [19] The various resonators and oscillator designs developed for dye lasers have been successfully adapted to other laser types such as the diode laser. [20] The physics of narrow-linewidth multiple-prism grating lasers was explained by Duarte and Piper. [21]

Chemicals used

Rhodamine 6G Chloride powder; mixed with methanol; emitting yellow light under the influence of a green laser Rhodamine6G Chloride.JPG
Rhodamine 6G Chloride powder; mixed with methanol; emitting yellow light under the influence of a green laser

Some of the laser dyes are rhodamine (orange, 540–680 nm), fluorescein (green, 530–560 nm), coumarin (blue 490–620 nm), stilbene (violet 410–480 nm), umbelliferone (blue, 450–470 nm), tetracene, malachite green, and others. [22] [23] While some dyes are actually used in food coloring, most dyes are very toxic, and often carcinogenic. [24] Many dyes, such as rhodamine 6G, (in its chloride form), can be very corrosive to all metals except stainless steel. Although dyes have very broad fluorescence spectra, the dye's absorption and emission will tend to center on a certain wavelength and taper off to each side, forming a tunability curve, with the absorption center being of a shorter wavelength than the emission center. Rhodamine 6G, for example, has its highest output around 590 nm, and the conversion efficiency lowers as the laser is tuned to either side of this wavelength.

A wide variety of solvents can be used, although most dyes will dissolve better in some solvents than in others. Some of the solvents used are water, glycol, ethanol, methanol, hexane, cyclohexane, cyclodextrin, and many others. Solvents can be highly toxic, and can sometimes be absorbed directly through the skin, or through inhaled vapors. Many solvents are also extremely flammable. The various solvents can also have an effect on the specific color of the dye solution, the lifetime of the singlet state, either enhancing or quenching the triplet state, and, thus, on the lasing bandwidth and power obtainable with a particular laser-pumping source. [10]

Adamantane is added to some dyes to prolong their life.

Cycloheptatriene and cyclooctatetraene (COT) can be added as triplet quenchers for rhodamine G, increasing the laser output power. Output power of 1.4 kilowatt at 585 nm was achieved using Rhodamine 6G with COT in methanol-water solution.

Excitation lasers

Flashlamps and several types of lasers can be used to optically pump dye lasers. A partial list of excitation lasers include: [25]

Ultra-short optical pulses

R. L. Fork, B. I. Greene, and C. V. Shank demonstrated, in 1981, the generation of ultra-short laser pulse using a ring-dye laser (or dye laser exploiting colliding pulse mode-locking). This kind of laser is capable of generating laser pulses of ~ 0.1 ps duration. [26]

The introduction of grating techniques and intra-cavity prismatic pulse compressors eventually resulted in the routine emission of femtosecond dye laser pulses.

Applications

An atomic vapor laser isotope separation experiment at LLNL. Green light is from a copper vapor pump laser used to pump a highly tuned dye laser which is producing the orange light. AVLIS laser.jpg
An atomic vapor laser isotope separation experiment at LLNL. Green light is from a copper vapor pump laser used to pump a highly tuned dye laser which is producing the orange light.

Dye lasers are very versatile. In addition to their recognized wavelength agility these lasers can offer very large pulsed energies or very high average powers. Flashlamp-pumped dye lasers have been shown to yield hundreds of Joules per pulse and copper-laser-pumped dye lasers are known to yield average powers in the kilowatt regime. [27]

Dye lasers are used in many applications including:

In laser medicine these lasers are applied in several areas, [31] [32] including dermatology where they are used to make skin tone more even. The wide range of wavelengths possible allows very close matching to the absorption lines of certain tissues, such as melanin or hemoglobin, while the narrow bandwidth obtainable helps reduce the possibility of damage to the surrounding tissue. They are used to treat port-wine stains and other blood vessel disorders, scars and kidney stones. They can be matched to a variety of inks for tattoo removal, as well as a number of other applications. [33]

In spectroscopy, dye lasers can be used to study the absorption and emission spectra of various materials. Their tunability, (from the near-infrared to the near-ultraviolet), narrow bandwidth, and high intensity allows a much greater diversity than other light sources. The variety of pulse widths, from ultra-short, femtosecond pulses to continuous-wave operation, makes them suitable for a wide range of applications, from the study of fluorescent lifetimes and semiconductor properties to lunar laser ranging experiments. [34]

Tunable lasers are used in swept-frequency metrology to enable measurement of absolute distances with very high accuracy. A two axis interferometer is set up and by sweeping the frequency, the frequency of the light returning from the fixed arm is slightly different from the frequency returning from the distance measuring arm. This produces a beat frequency which can be detected and used to determine the absolute difference between the lengths of the two arms. [35]

See also

Related Research Articles

<span class="mw-page-title-main">Laser</span> Device which emits light via optical amplification

A laser is a device that emits light through a process of optical amplification based on the stimulated emission of electromagnetic radiation. The word laser is an anacronym that originated as an acronym for light amplification by stimulated emission of radiation. The first laser was built in 1960 by Theodore Maiman at Hughes Research Laboratories, based on theoretical work by Charles H. Townes and Arthur Leonard Schawlow.

<span class="mw-page-title-main">Laser construction</span>

A laser is constructed from three principal parts:

<span class="mw-page-title-main">Laser diode</span> Semiconductor laser

A laser diode is a semiconductor device similar to a light-emitting diode in which a diode pumped directly with electrical current can create lasing conditions at the diode's junction.

<span class="mw-page-title-main">Flashtube</span> Incoherent light source

A flashtube (flashlamp) is an electric arc lamp designed to produce extremely intense, incoherent, full-spectrum white light for a very short time. A flashtube is a glass tube with an electrode at each end and is filled with a gas that, when triggered, ionizes and conducts a high-voltage pulse to make light. Flashtubes are used most in photography; they also are used in science, medicine, industry, and entertainment.

<span class="mw-page-title-main">Nd:YAG laser</span> Crystal used as a lasing medium for solid-state lasers

Nd:YAG (neodymium-doped yttrium aluminum garnet; Nd:Y3Al5O12) is a crystal that is used as a lasing medium for solid-state lasers. The dopant, triply ionized neodymium, Nd(III), typically replaces a small fraction (1%) of the yttrium ions in the host crystal structure of the yttrium aluminum garnet (YAG), since the two ions are of similar size. It is the neodymium ion which provides the lasing activity in the crystal, in the same fashion as red chromium ion in ruby lasers.

<span class="mw-page-title-main">Excimer laser</span> Type of ultraviolet laser important in chip manufacturing and eye surgery

An excimer laser, sometimes more correctly called an exciplex laser, is a form of ultraviolet laser which is commonly used in the production of microelectronic devices, semiconductor based integrated circuits or "chips", eye surgery, and micromachining. Since 1960s excimer lasers are widely used in high-resolution photolithography machines, one of the critical technologies required for microelectronic chip manufacturing.

<span class="mw-page-title-main">Gas laser</span> Laser in which electricity is discharged through gas

A gas laser is a laser in which an electric current is discharged through a gas to produce coherent light. The gas laser was the first continuous-light laser and the first laser to operate on the principle of converting electrical energy to a laser light output. The first gas laser, the Helium–neon laser (HeNe), was co-invented by Iranian engineer and scientist Ali Javan and American physicist William R. Bennett, Jr., in 1960. It produced a coherent light beam in the infrared region of the spectrum at 1.15 micrometres.

<span class="mw-page-title-main">Tunable laser</span>

A tunable laser is a laser whose wavelength of operation can be altered in a controlled manner. While all laser gain media allow small shifts in output wavelength, only a few types of lasers allow continuous tuning over a significant wavelength range.

<span class="mw-page-title-main">Rhodamine 6G</span> Chemical compound

Rhodamine 6G is a highly fluorescent rhodamine family dye. It is often used as a tracer dye within water to determine the rate and direction of flow and transport. Rhodamine dyes fluoresce and can thus be detected easily and inexpensively with instruments called fluorometers. Rhodamine dyes are used extensively in biotechnology applications such as fluorescence microscopy, flow cytometry, fluorescence correlation spectroscopy and ELISA.

<span class="mw-page-title-main">Solid-state laser</span> Laser which uses a solid gain medium

A solid-state laser is a laser that uses a gain medium that is a solid, rather than a liquid as in dye lasers or a gas as in gas lasers. Semiconductor-based lasers are also in the solid state, but are generally considered as a separate class from solid-state lasers, called laser diodes.

<span class="mw-page-title-main">Laser pumping</span> Powering mechanism for lasers

Laser pumping is the act of energy transfer from an external source into the gain medium of a laser. The energy is absorbed in the medium, producing excited states in its atoms. When for a period of time the number of particles in one excited state exceeds the number of particles in the ground state or a less-excited state, population inversion is achieved. In this condition, the mechanism of stimulated emission can take place and the medium can act as a laser or an optical amplifier. The pump power must be higher than the lasing threshold of the laser.

Ultrafast laser spectroscopy is a spectroscopic technique that uses ultrashort pulse lasers for the study of dynamics on extremely short time scales. Different methods are used to examine the dynamics of charge carriers, atoms, and molecules. Many different procedures have been developed spanning different time scales and photon energy ranges; some common methods are listed below.

<span class="mw-page-title-main">Ruby laser</span> Solid-state laser

A ruby laser is a solid-state laser that uses a synthetic ruby crystal as its gain medium. The first working laser was a ruby laser made by Theodore H. "Ted" Maiman at Hughes Research Laboratories on May 16, 1960.

<span class="mw-page-title-main">Prism compressor</span>

A prism compressor is an optical device used to shorten the duration of a positively chirped ultrashort laser pulse by giving different wavelength components a different time delay. It typically consists of two prisms and a mirror. Figure 1 shows the construction of such a compressor. Although the dispersion of the prism material causes different wavelength components to travel along different paths, the compressor is built such that all wavelength components leave the compressor at different times, but in the same direction. If the different wavelength components of a laser pulse were already separated in time, the prism compressor can make them overlap with each other, thus causing a shorter pulse.

<span class="mw-page-title-main">Output coupler</span> Part of an optical resonator which allows intracavity light to be emitted

In laser science, an output coupler (OC) is the component of an optical resonator that allows the extraction of a portion of the light from the laser's intracavity beam. An output coupler most often consists of a partially reflective mirror, allowing a certain portion of the intracavity beam to transmit through. Other methods include the use of almost-totally reflective mirrors at each end of the cavity, emitting the beam either by focusing it into a small hole drilled in the center of one mirror, or by redirecting through the use of rotating mirrors, prisms, or other optical devices, causing the beam to bypass one of the end mirrors at a given time.

<span class="mw-page-title-main">F. J. Duarte</span>

Francisco Javier "Frank" Duarte is a laser physicist and author/editor of several books on tunable lasers.

Beam expanders are optical devices that take a collimated beam of light and expand its width.

<span class="mw-page-title-main">Solid-state dye laser</span>

A solid-state dye laser (SSDL) is a solid-state lasers in which the gain medium is a laser dye-doped organic matrix such as poly(methyl methacrylate) (PMMA), rather than a liquid solution of the dye. These lasers are also referred to as solid-state organic lasers and solid-state dye-doped polymer lasers.

<span class="mw-page-title-main">Multiple-prism grating laser oscillator</span>

Multiple-prism grating laser oscillators, or MPG laser oscillators, use multiple-prism beam expansion to illuminate a diffraction grating mounted either in Littrow configuration or grazing-incidence configuration. Originally, these narrow-linewidth tunable dispersive oscillators were introduced as multiple-prism Littrow (MPL) grating oscillators, or hybrid multiple-prism near-grazing-incidence (HMPGI) grating cavities, in organic dye lasers. However, these designs were quickly adopted for other types of lasers such as gas lasers, diode lasers, and more recently fiber lasers.

<span class="mw-page-title-main">Organic laser</span> Laser that uses a carbon-based material as the gain medium

An organic laser is a laser which uses an organic material as the gain medium. The first organic laser was the liquid dye laser. These lasers use laser dye solutions as their gain media.

References

  1. Dye Laser Principles: With Applications by Frank J. Duarte, Lloyd W. Hillman -- Academic Press 1990 Page 42
  2. F. P. Schäfer (Ed.), Dye Lasers (Springer-Verlag, Berlin, 1990).
  3. F. J. Duarte and L. W. Hillman (Eds.), Dye Laser Principles (Academic, New York, 1990).
  4. Design and Analysis of Flashlamp Systems for Pumping Organic Dye Lasers – J. F. Holzrichter and A. L. Schawlow. Annals of the New York Academy of Sciences
  5. Yee, T. K.; Fan, B.; Gustafson, T. K. (1979-04-15). "Simmer-enhanced flashlamp-pumped dye laser". Applied Optics. The Optical Society. 18 (8): 1131–2. Bibcode:1979ApOpt..18.1131Y. doi:10.1364/ao.18.001131. ISSN   0003-6935. PMID   20208893.
  6. "General Xenon Flash and Strobe Design Guidelines". members.misty.com. Retrieved 19 April 2018.
  7. "Sam's Laser FAQ - Home-Built Dye Laser". www.repairfaq.org. Retrieved 19 April 2018.
  8. Paschotta, Dr. Rüdiger. "Encyclopedia of Laser Physics and Technology - spatial hole burning, SHB, laser, single-frequency operation". www.rp-photonics.com. Retrieved 19 April 2018.
  9. Laser fundamentals by William T. Silfvast – Cambridge University Press 1996 Page 397-399
  10. 1 2 3 "Archived copy" (PDF). Archived from the original (PDF) on 2017-02-16. Retrieved 2017-02-13.{{cite web}}: CS1 maint: archived copy as title (link)
  11. 1 2 "Principles of Lasers", by Orazio Svelto
  12. F. P. Schäfer and K. H. Drexhage, Dye Lasers., 2nd rev. ed., vol. 1, Berlin ; New York: Springer-Verlag, 1977
  13. O. G. Peterson, S. A. Tuccio, B. B. Snavely, "CW operation of an organic dye solution laser", Appl. Phys. Lett.42, 1917-1918 (1970).
  14. F. J. Duarte and L. W. Hillman, Dye Laser Principles (Academic, New York, 1990) Chapter 4.
  15. T. W. Hänsch, Repetitively Pulsed Tunable Dye Laser for High Resolution Spectroscopy, Appl. Opt.11, 895-898 (1972).
  16. I. Shoshan, N. N. Danon, and U. P. Oppenheim, Narrowband operation of a pulsed dye laser without intracavity beam expansion, J. Appl. Phys.48, 4495-4497 (1977).
  17. Littman, Michael G.; Metcalf, Harold J. (1978-07-15). "Spectrally narrow pulsed dye laser without beam expander". Applied Optics. The Optical Society. 17 (14): 2224–7. Bibcode:1978ApOpt..17.2224L. doi:10.1364/ao.17.002224. ISSN   0003-6935. PMID   20203761.
  18. Duarte, F.J.; Piper, J.A. (1980). "A double-prism beam expander for pulsed dye lasers". Optics Communications. Elsevier BV. 35 (1): 100–104. Bibcode:1980OptCo..35..100D. doi:10.1016/0030-4018(80)90368-5. ISSN   0030-4018.
  19. Duarte, F. J.; Piper, J. A. (1981-06-15). "Prism preexpanded grazing-incidence grating cavity for pulsed dye lasers". Applied Optics. The Optical Society. 20 (12): 2113–6. Bibcode:1981ApOpt..20.2113D. doi:10.1364/ao.20.002113. ISSN   0003-6935. PMID   20332895.
  20. P. Zorabedian, Tunable external cavity semiconductor lasers, in Tunable Lasers Handbook, F. J. Duarte (Ed.) (Academic, New York, 1995) Chapter 8.
  21. Duarte, F.J.; Piper, J.A. (1982). "Dispersion theory of multiple-prism beam expanders for pulsed dye lasers". Optics Communications. Elsevier BV. 43 (5): 303–307. Bibcode:1982OptCo..43..303D. doi:10.1016/0030-4018(82)90216-4. ISSN   0030-4018.
  22. Amnon Yariv, Optical Electronics in Modern Communications, Fifth Edition, page 266
  23. "Tuning Curves" (PDF). Exciton. Archived from the original (PDF) on 2011-09-20. Retrieved 2023-11-03.
  24. "Archived copy" (PDF). Archived from the original (PDF) on 2015-02-21. Retrieved 2012-08-15.{{cite web}}: CS1 maint: archived copy as title (link)
  25. F. J. Duarte and L. W. Hillman (Eds.), Dye Laser Principles (Academic, New York, 1990) Chapters 5 and 6.
  26. Fork, R. L.; Greene, B. I.; Shank, C. V. (1981). "Generation of optical pulses shorter than 0.1 psec by colliding pulse mode locking". Applied Physics Letters. AIP Publishing. 38 (9): 671–672. Bibcode:1981ApPhL..38..671F. doi:10.1063/1.92500. ISSN   0003-6951. S2CID   45813878.
  27. "HIGH POWER DYE LASERS". www.tunablelasers.com. Retrieved 19 April 2018.
  28. M. A. Akerman, Dye laser isotope separation, in Dye Laser Principles, F. J. Duarte and L. W. Hillman (eds.)(Academic, New York, 1990) Chapter 9.
  29. D. Klick, Industrial applications of dye lasers, in Dye Laser Principles, F. J. Duarte and L. W. Hillman (eds.)(Academic, New York, 1990) Chapter 8.
  30. W. Demtröder, Laser Spectroscopy, 3rd Ed. (Springer, 2003).
  31. L. Goldman, Dye lasers in medicine, in Dye Laser Principles, F. J. Duarte and L. W. Hillman, Eds. (Academic, New York, 1990) Chapter 10.
  32. Costela A, Garcia-Moreno I, Gomez C (2016). "Medical Applications of Organic Dye Lasers". In Duarte FJ (ed.). Tunable Laser Applications (3rd ed.). Boca Raton: CRC Press. pp. 293–313. ISBN   9781482261066.
  33. Duarte FJ, ed. (2016). Tunable Laser Applications (3rd ed.). Boca Raton: CRC Press. ISBN   9781482261066.
  34. The Laser Guidebook By Jeff Hecht – McGraw Hill 1992 Page 294
  35. "Highly linear, Widerange Swept Frequency Generation at Microwave and Optical Frequencies" (PDF). nasa.gov. Archived from the original (PDF) on 7 September 2012. Retrieved 19 April 2018.