Laser linewidth

Last updated

Laser linewidth is the spectral linewidth of a laser beam.

Contents

Two of the most distinctive characteristics of laser emission are spatial coherence and spectral coherence. While spatial coherence is related to the beam divergence of the laser, spectral coherence is evaluated by measuring the linewidth of laser radiation.

Theory

History: First derivation of the laser linewidth

The first human-made coherent light source was a maser. The acronym MASER stands for "Microwave Amplification by Stimulated Emission of Radiation". More precisely, it was the ammonia maser operating at 12.5 mm wavelength that was demonstrated by Gordon, Zeiger, and Townes in 1954. [1] One year later the same authors derived [2] theoretically the linewidth of their device by making the reasonable approximations that their ammonia maser

  1. is a true continuous-wave (CW) maser, [2]
  2. is a true four-level maser, [2] and
  3. exhibits no intrinsic resonator losses but only outcoupling losses. [2]

Notably, their derivation was entirely semi-classical, [2] describing the ammonia molecules as quantum emitters and assuming classical electromagnetic fields (but no quantized fields or quantum fluctuations), resulting in the half-width-at-half-maximum (HWHM) maser linewidth [2]

denoted here by an asterisk and converted to the full-width-at-half-maximum (FWHM) linewidth . is the Boltzmann constant, is the temperature, is the output power, and and are the HWHM and FWHM linewidths of the underlying passive microwave resonator, respectively.

In 1958, two years before Maiman demonstrated the laser (initially called an "optical maser"), [3] Schawlow and Townes [4] transferred the maser linewidth to the optical regime by replacing the thermal energy by the photon energy , where is the Planck constant and is the frequency of laser light, thereby approximating that

iv. one photon is coupled into the lasing mode by spontaneous emission during the photon-decay time , [5]

resulting in the original Schawlow–Townes approximation of the laser linewidth: [4]

Again, the transfer from the microwave to the optical regime was entirely semi-classical. Consequently, the original Schawlow–Townes equation is entirely based on semi-classical physics [2] [4] and is a four-fold approximation of a more general laser linewidth, [5] which will be derived in the following.

Passive resonator mode: Photon-decay time

We assume a two-mirror Fabry–Pérot resonator [6] of geometrical length , homogeneously filled with an active laser medium of refractive index . We define the reference situation, namely the passive resonator mode, for a resonator whose active medium is transparent, i.e., it does not introduce gain or absorption.

The round-trip time of light travelling in the resonator with speed , where is the speed of light in vacuum, and the free spectral range are given by [6] [5]

Light in the longitudinal resonator mode of interest oscillates at the qth resonance frequency [6] [5]

The exponential outcoupling decay time and the corresponding decay-rate constant are related to the intensity reflectances of the two resonator mirrors by [6] [5]

The exponential intrinsic loss time and the corresponding decay-rate constant are related to the intrinsic round-trip loss by [5]

The exponential photon-decay time and the corresponding decay-rate constant of the passive resonator are then given by [5]

All three exponential decay times average over the round-trip time [5] In the following, we assume that , , , , and , hence also , , and do not vary significantly over the frequency range of interest.

Passive resonator mode: Lorentzian linewidth, Q-factor, coherence time and length

Besides the photon-decay time , the spectral-coherence properties of the passive resonator mode can be equivalently expressed by the following parameters. The FWHM Lorentzian linewidth of the passive resonator mode that appears in the Schawlow–Townes equation is derived from the exponential photon-decay time by Fourier transformation, [6] [5]

The Q-factor is defined as the energy stored in the resonator mode over the energy lost per oscillation cycle, [5]

where is the number of photons in the mode. The coherence time and coherence length of light emitted from the mode are given by [5]

Active resonator mode: Gain, photon-decay time, Lorentzian linewidth, Q-factor, coherence time and length

With the population densities and of upper and lower laser level, respectively, and the effective cross sections and of stimulated emission and absorption at the resonance frequency , respectively, the gain per unit length in the active laser medium at the resonance frequency is given by [5]

A value of induces amplification, whereas induces absorption of light at the resonance frequency , resulting in an elongated or shortened photon-decay time of photons out of the active resonator mode, respectively, [5]

The other four spectral-coherence properties of the active resonator mode are obtained in the same way as for the passive resonator mode. The Lorentzian linewidth is derived by Fourier transformation, [5]

A value of leads to gain narrowing, whereas leads to absorption broadening of the spectral linewidth. The Q-factor is [5]

The coherence time and length are [5]

Spectral-coherence factor

The factor by which the photon-decay time is elongated by gain or shortened by absorption is introduced here as the spectral-coherence factor : [5]

All five spectral-coherence parameters then scale by the same spectral-coherence factor : [5]

Lasing resonator mode: Fundamental laser linewidth

With the number of photons propagating inside the lasing resonator mode, the stimulated-emission and photon-decay rates are, respectively, [5]

The spectral-coherence factor then becomes [5]

The photon-decay time of the lasing resonator mode is [5]

The fundamental laser linewidth is [5]

This fundamental linewidth is valid for lasers with an arbitrary energy-level system, operating below, at, or above threshold, with the gain being smaller, equal, or larger compared to the losses, and in a cw or a transient lasing regime. [5]

It becomes clear from its derivation that the fundamental laser linewidth is due to the semi-classical effect that the gain elongates the photon-decay time. [5]

Continuous-wave laser: The gain is smaller than the losses

The spontaneous-emission rate into the lasing resonator mode is given by [5]

Notably, is always a positive rate, because one atomic excitation is converted into one photon in the lasing mode. [7] [5] It is the source term of laser radiation and must not be misinterpreted as "noise". [5] The photon-rate equation for a single lasing mode reads [5]

A CW laser is defined by a temporally constant number of photons in the lasing mode, hence . In a CW laser the stimulated- and spontaneous-emission rates together compensate the photon-decay rate. Consequently, [5]

The stimulated-emission rate is smaller than the photon-decay rate or, colloquially, "the gain is smaller than the losses". [5] This fact has been known for decades and exploited to quantify the threshold behavior of semiconductor lasers. [8] [9] [10] [11] Even far above laser threshold the gain is still a tiny bit smaller than the losses. It is exactly this small difference that induces the finite linewidth of a CW laser. [5]

It becomes clear from this derivation that fundamentally the laser is an amplifier of spontaneous emission, and the cw laser linewidth is due to the semi-classical effect that the gain is smaller than the losses. [5] Also in the quantum-optical approaches to the laser linewidth, [12] based on the density-operator master equation, it can be verified that the gain is smaller than the losses. [5]

Schawlow–Townes approximation

As mentioned above, it is clear from its historical derivation that the original Schawlow–Townes equation is a four-fold approximation of the fundamental laser linewidth. Starting from the fundamental laser linewidth derived above, by applying the four approximations i.–iv. one then obtains the original Schawlow–Townes equation.

  1. It is a true CW laser, hence [5]
  2. It is a true four-level laser, hence [5]
  3. It has no intrinsic resonator losses, hence [5]
  4. One photon is coupled into the lasing mode by spontaneous emission during the photon-decay time , which would happen exactly at the unreachable point of an ideal four-level CW laser with infinite spectral-coherence factor , photon number , and output power , where the gain would equal the losses, hence [5]

I.e., by applying the same four approximations i.–iv. to the fundamental laser linewidth that were applied in the first derivation, [2] [4] the original Schawlow–Townes equation is obtained. [5]

Thus, the fundamental laser linewidth is [5]

whereas the original Schawlow–Townes equation is a four-fold approximation of this fundamental laser linewidth and is merely of historical interest.

Additional linewidth broadening and narrowing effects

Following its publication in 1958, [4] the original Schawlow–Townes equation was extended in various ways. These extended equations often trade under the same name, the "Schawlow–Townes linewidth", thereby creating a veritable confusion in the available literature on the laser linewidth, as it is often unclear which particular extension of the original Schawlow–Townes equation the respective authors refer to.

Several semi-classical extensions intended to remove one or several of the approximations i.–iv. mentioned above, thereby making steps towards the fundamental laser linewidth derived above.

The following extensions may add to the fundamental laser linewidth:

  1. Hempstead and Lax, [13] as well as Haken, [14] predicted quantum-mechanically an additional linewidth narrowing by a factor of two near laser threshold. However, such an effect was observed experimentally only in a handful of cases.
  2. Petermann derived semi-classically a previously experimentally observed linewidth-broadening effect in gain-guided compared to index-guided semiconductor waveguide lasers. [15] Siegman later showed that this effect is due to the non-orthogonality of transverse modes. [16] [17] Woerdman and co-workers extended this idea to longitudinal modes [18] and polarization modes. [19] As a result, the so-called "Petermann K-factor" is sometimes added to the laser linewidth.
  3. Henry predicted quantum-mechanically an additional linewidth broadening due to refractive-index changes related to electron-hole-pair excitation, which induce phase changes. [20] As a result, the so-called "Henry's -factor" is sometimes added to the laser linewidth.

Measurement of laser linewidth

One of the first methods used to measure the coherence of a laser was interferometry. [21] A typical method to measure the laser linewidth is self-heterodyne interferometry. [22] [23] An alternative approach is the use of spectrometry. [24]

Continuous lasers

The laser linewidth in a typical single-transverse-mode He–Ne laser (at a wavelength of 632.8 nm), in the absence of intracavity line narrowing optics, can be on the order of 1 GHz. Rare-earth-doped dielectric-based or semiconductor-based distributed-feedback lasers have typical linewidths on the order of 1 kHz. [25] [26] The laser linewidth from stabilized low-power continuous-wave lasers can be very narrow and reach down to less than 1 kHz. [27] Observed linewidths are larger than the fundamental laser linewidth due to technical noise (temporal fluctuations of the optical pump power or pump current, mechanical vibrations, refractive-index and length changes due to temperature fluctuations, etc.).

Pulsed lasers

Laser linewidth from high-power, high-gain pulsed-lasers, in the absence of intracavity line narrowing optics, can be quite broad and in the case of powerful broadband dye lasers it can range from a few nm wide [28] to as broad as 10 nm. [24]

Laser linewidth from high-power high-gain pulsed laser oscillators, comprising line narrowing optics, is a function of the geometrical and dispersive features of the laser cavity. [29] To a first approximation the laser linewidth, in an optimized cavity, is directly proportional to the beam divergence of the emission multiplied by the inverse of the overall intracavity dispersion. [29] That is,

This is known as the cavity linewidth equation where is the beam divergence and the term in parentheses (elevated to −1) is the overall intracavity dispersion. This equation was originally derived from classical optics. [30] However, in 1992 Duarte derived this equation from quantum interferometric principles, [31] thus linking a quantum expression with the overall intracavity angular dispersion.

An optimized multiple-prism grating laser oscillator can deliver pulse emission in the kW regime at single-longitudinal-mode linewidths of ≈ 350 MHz (equivalent to ≈ 0.0004 nm at a laser wavelength of 590 nm). [32] Since the pulse duration from these oscillators is about 3 ns, [32] the laser linewidth performance is near the limit allowed by the Heisenberg uncertainty principle.

See also

Related Research Articles

In physics, specifically statistical mechanics, a population inversion occurs while a system exists in a state in which more members of the system are in higher, excited states than in lower, unexcited energy states. It is called an "inversion" because in many familiar and commonly encountered physical systems, this is not possible. This concept is of fundamental importance in laser science because the production of a population inversion is a necessary step in the workings of a standard laser.

<span class="mw-page-title-main">Stimulated emission</span> Release of a photon triggered by another

Stimulated emission is the process by which an incoming photon of a specific frequency can interact with an excited atomic electron, causing it to drop to a lower energy level. The liberated energy transfers to the electromagnetic field, creating a new photon with a frequency, polarization, and direction of travel that are all identical to the photons of the incident wave. This is in contrast to spontaneous emission, which occurs at a characteristic rate for each of the atoms/oscillators in the upper energy state regardless of the external electromagnetic field.

In physics, coherence length is the propagation distance over which a coherent wave maintains a specified degree of coherence. Wave interference is strong when the paths taken by all of the interfering waves differ by less than the coherence length. A wave with a longer coherence length is closer to a perfect sinusoidal wave. Coherence length is important in holography and telecommunications engineering.

For an electromagnetic wave, the coherence time is the time over which a propagating wave may be considered coherent, meaning that its phase is, on average, predictable.

Brightness temperature or radiance temperature is a measure of the intensity of electromagnetic energy coming from a source. In particular, it is the temperature at which a black body would have to be in order to duplicate the observed intensity of a grey body object at a frequency . This concept is used in radio astronomy, planetary science, materials science and climatology.

<span class="mw-page-title-main">Noether's theorem</span> Statement relating differentiable symmetries to conserved quantities

Noether's theorem states that every continuous symmetry of the action of a physical system with conservative forces has a corresponding conservation law. This is the first of two theorems proven by mathematician Emmy Noether in 1915 and published in 1918. The action of a physical system is the integral over time of a Lagrangian function, from which the system's behavior can be determined by the principle of least action. This theorem only applies to continuous and smooth symmetries of physical space.

<span class="mw-page-title-main">Fabry–Pérot interferometer</span> Optical device with parallel mirrors

In optics, a Fabry–Pérot interferometer (FPI) or etalon is an optical cavity made from two parallel reflecting surfaces. Optical waves can pass through the optical cavity only when they are in resonance with it. It is named after Charles Fabry and Alfred Perot, who developed the instrument in 1899. Etalon is from the French étalon, meaning "measuring gauge" or "standard".

<i>Q</i> factor Parameter describing the longevity of energy in a resonator relative to its resonant frequency

In physics and engineering, the quality factor or Q factor is a dimensionless parameter that describes how underdamped an oscillator or resonator is. It is defined as the ratio of the initial energy stored in the resonator to the energy lost in one radian of the cycle of oscillation. Q factor is alternatively defined as the ratio of a resonator's centre frequency to its bandwidth when subject to an oscillating driving force. These two definitions give numerically similar, but not identical, results. Higher Q indicates a lower rate of energy loss and the oscillations die out more slowly. A pendulum suspended from a high-quality bearing, oscillating in air, has a high Q, while a pendulum immersed in oil has a low one. Resonators with high quality factors have low damping, so that they ring or vibrate longer.

The laser diode rate equations model the electrical and optical performance of a laser diode. This system of ordinary differential equations relates the number or density of photons and charge carriers (electrons) in the device to the injection current and to device and material parameters such as carrier lifetime, photon lifetime, and the optical gain.

In physics and astronomy, the Reissner–Nordström metric is a static solution to the Einstein–Maxwell field equations, which corresponds to the gravitational field of a charged, non-rotating, spherically symmetric body of mass M. The analogous solution for a charged, rotating body is given by the Kerr–Newman metric.

In general relativity, Schwarzschild geodesics describe the motion of test particles in the gravitational field of a central fixed mass that is, motion in the Schwarzschild metric. Schwarzschild geodesics have been pivotal in the validation of Einstein's theory of general relativity. For example, they provide accurate predictions of the anomalous precession of the planets in the Solar System and of the deflection of light by gravity.

Quantum noise is noise arising from the indeterminate state of matter in accordance with fundamental principles of quantum mechanics, specifically the uncertainty principle and via zero-point energy fluctuations. Quantum noise is due to the apparently discrete nature of the small quantum constituents such as electrons, as well as the discrete nature of quantum effects, such as photocurrents.

<span class="mw-page-title-main">Longitudinal mode</span> Standing wave patterns of resonator cavities

A longitudinal mode of a resonant cavity is a particular standing wave pattern formed by waves confined in the cavity. The longitudinal modes correspond to the wavelengths of the wave which are reinforced by constructive interference after many reflections from the cavity's reflecting surfaces. All other wavelengths are suppressed by destructive interference.

<span class="mw-page-title-main">Jaynes–Cummings model</span> Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

Free spectral range (FSR) is the spacing in optical frequency or wavelength between two successive reflected or transmitted optical intensity maxima or minima of an interferometer or diffractive optical element.

The Purcell effect is the enhancement of a quantum system's spontaneous emission rate by its environment. In the 1940s Edward Mills Purcell discovered the enhancement of spontaneous emission rates of atoms when they are incorporated into a resonant cavity. In terms of quantum electrodynamics the Purcell effect is a consequence of enhancement of local density of photonic states at the emitter position. It can also be considered as an interference effect. The oscillator radiates the wave which is reflected from the environment. In turn the reflection excites the oscillator either out of phase resulting in higher damping rate accompanied with the radiation enhancement or in phase with the oscillator mode leading to the radiation suppression.

Circuit quantum electrodynamics provides a means of studying the fundamental interaction between light and matter. As in the field of cavity quantum electrodynamics, a single photon within a single mode cavity coherently couples to a quantum object (atom). In contrast to cavity QED, the photon is stored in a one-dimensional on-chip resonator and the quantum object is no natural atom but an artificial one. These artificial atoms usually are mesoscopic devices which exhibit an atom-like energy spectrum. The field of circuit QED is a prominent example for quantum information processing and a promising candidate for future quantum computation.

<span class="mw-page-title-main">White light interferometry</span> Measurement technique

As described here, white light interferometry is a non-contact optical method for surface height measurement on 3D structures with surface profiles varying between tens of nanometers and a few centimeters. It is often used as an alternative name for coherence scanning interferometry in the context of areal surface topography instrumentation that relies on spectrally-broadband, visible-wavelength light.

<span class="mw-page-title-main">Cavity optomechanics</span>

Cavity optomechanics is a branch of physics which focuses on the interaction between light and mechanical objects on low-energy scales. It is a cross field of optics, quantum optics, solid-state physics and materials science. The motivation for research on cavity optomechanics comes from fundamental effects of quantum theory and gravity, as well as technological applications.

Laser theory of Fabry-Perot (FP) semiconductor lasers proves to be nonlinear, since the gain, the refractive index and the loss coefficient are the functions of energy flux. The nonlinear theory made it possible to explain a number of experiments some of which could not even be explained, much less modeled, on the basis of other theoretical models; this suggests that the nonlinear theory developed is a new paradigm of the laser theory.

References

  1. Gordon, J. P.; Zeiger, H. J.; Townes, C. H. (1954). "Molecular microwave oscillator and new hyperfine structure in the microwave spectrum of NH3". Physical Review. 95 (1): 282–284. Bibcode:1954PhRv...95..282G. doi: 10.1103/PhysRev.95.282 .
  2. 1 2 3 4 5 6 7 8 Gordon, J. P.; Zeiger, H. J.; Townes, C. H. (1955). "The maser−New type of microwave amplifier, frequency standard, and spectrometer". Physical Review. 99 (4): 1264–1274. Bibcode:1955PhRv...99.1264G. doi: 10.1103/PhysRev.99.1264 .
  3. Maiman, T. H. (1960). "Stimulated optical radiation in Ruby". Nature. 187 (4736): 493–494. Bibcode:1960Natur.187..493M. doi:10.1038/187493a0. S2CID   4224209.
  4. 1 2 3 4 5 Schawlow, A. L.; Townes, C. H. (1958). "Infrared and optical masers". Physical Review. 112 (6): 1940–1949. Bibcode:1958PhRv..112.1940S. doi: 10.1103/PhysRev.112.1940 .
  5. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 Pollnau, M.; Eichhorn, M. (2020). "Spectral coherence, Part I: Passive resonator linewidth, fundamental laser linewidth, and Schawlow–Townes approximation". Progress in Quantum Electronics. 72: 100255. Bibcode:2020PQE....7200255P. doi: 10.1016/j.pquantelec.2020.100255 .
  6. 1 2 3 4 5 Ismail, N.; Kores, C. C.; Geskus, D.; Pollnau, M. (2016). "Fabry–Pérot resonator: spectral line shapes, generic and related Airy distributions, linewidths, finesses, and performance at low or frequency-dependent reflectivity" (PDF). Optics Express. 24 (15): 16366–16389. Bibcode:2016OExpr..2416366I. doi: 10.1364/OE.24.016366 . PMID   27464090.
  7. Pollnau, M. (2018). "Phase aspect in photon emission and absorption" (PDF). Optica. 5 (4): 465–474. Bibcode:2018Optic...5..465P. doi: 10.1364/OPTICA.5.000465 .
  8. Sommers, H. S. (1974). "Spontaneous power and the coherent state of injection lasers". Journal of Applied Physics. 45 (4): 1787–1793. Bibcode:1974JAP....45.1787S. doi:10.1063/1.1663491.
  9. Sommers, H. S. (1982). "Threshold and oscillation of injection lasers: a critical review of laser theory". Solid-State Electronics. 25 (1): 25–44. Bibcode:1982SSEle..25...25S. doi:10.1016/0038-1101(82)90091-0.
  10. Siegman, A. E. (1986) "Lasers", University Science Books, Mill Valley, California, ch. 13, pp. 510-524.
  11. Björk, G.; Yamamoto, Y. (1991). "Analysis of semiconductor microcavity lasers using rate equations". IEEE Journal of Quantum Electronics. 27 (11): 2386–2396. Bibcode:1991IJQE...27.2386B. doi:10.1109/3.100877.
  12. Sargent III, M.; Scully, M. O.; Lamb, Jr., W. E. (1993) "Laser Physics", 6th edition, Westview Press, Ch. 17.
  13. Hempstead, R. D.; Lax, M. (1967). "Classical noise. VI. Noise in self-sustained oscillators near threshold". Physical Review. 161 (2): 350–366. Bibcode:1967PhRv..161..350H. doi:10.1103/PhysRev.161.350.
  14. Haken, H. (1970) "Laser Theory", Vol. XXV/2c of Encyclopedia of Physics, Springer.
  15. Petermann, K. (1979). "Calculated spontaneous emission factor for double-heterostructure injection lasers with gain-induced waveguiding". IEEE Journal of Quantum Electronics. QE-15 (7): 566–570. Bibcode:1979IJQE...15..566P. doi:10.1109/JQE.1979.1070064.
  16. Siegman, A. E. (1989). "Excess spontaneous emission in non-Hermitian optical systems. I. Laser amplifiers". Physical Review A. 39 (3): 1253–1263. Bibcode:1989PhRvA..39.1253S. doi:10.1103/PhysRevA.39.1253. PMID   9901361.
  17. Siegman, A. E. (1989). "Excess spontaneous emission in non-Hermitian optical systems. II. Laser oscillators". Physical Review A. 39 (3): 1264–1268. Bibcode:1989PhRvA..39.1264S. doi:10.1103/PhysRevA.39.1264. PMID   9901362.
  18. Hamel, W. A.; Woerdman, J. P. (1989). "Nonorthogonality of the longitudinal eigenmodes of a laser". Physical Review A. 40 (5): 2785–2787. Bibcode:1989PhRvA..40.2785H. doi:10.1103/PhysRevA.40.2785. PMID   9902474.
  19. van der Lee, A. M.; van Druten, N. J.; Mieremet, A. L.; van Eijkelenborg, M. A.; Lindberg, Å. M.; van Exter, M. P.; Woerdman, J. P. (1989). "Excess quantum noise due to nonorthogonal polarization modes". Physical Review Letters. 79 (5): 4357–4360. Bibcode:1989PhRvA..40.2785H. doi:10.1103/PhysRevA.40.2785. PMID   9902474.
  20. Henry, C. H. (1982). "Theory of the linewidth of semiconductor lasers". IEEE Journal of Quantum Electronics. 18 (2): 259–264. Bibcode:1982IJQE...18..259H. doi:10.1109/JQE.1982.1071522.
  21. O. S. Heavens, Optical Masers (Wiley, New York, 1963).
  22. Okoshi, T.; Kikuchi, K.; Nakayama, A. (1980). "Novel method for high resolution measurement of laser output spectrum". Electronics Letters. 16 (16): 630–631. Bibcode:1980ElL....16..630O. doi:10.1049/el:19800437.
  23. Dawson, J. W.; Park, N.; Vahala, K. J. (1992). "An improved delayed self-heterodyne interferometer for linewidth measurements". IEEE Photonics Technology Letters. 4 (9): 1063–1066. Bibcode:1992IPTL....4.1063D. doi:10.1109/68.157150. S2CID   15033723.
  24. 1 2 Schäfer, Fritz P.; Schmidt, Werner; Volze, Jürgen (1966-10-15). "Organic Dye Solution Laser". Applied Physics Letters. 9 (8). AIP Publishing: 306–309. Bibcode:1966ApPhL...9..306S. doi: 10.1063/1.1754762 . ISSN   0003-6951.
  25. Bernhardi, E. H.; van Wolferen, H. A. G. M.; Agazzi, L.; Khan, M. R. H.; Roeloffzen, C. G. H.; Wörhoff, K.; Pollnau, M.; de Ridder, R. M. (2010). "Ultra-narrow-linewidth, single-frequency distributed feedback waveguide laser in Al2O3:Er3+ on silicon" (PDF). Optics Letters. 35 (14): 2394–2396. Bibcode:2010OptL...35.2394B. doi:10.1364/OL.35.002394. PMID   20634841.
  26. Santis, C. T.; Steger, S. T.; Vilenchik, Y.; Vasilyev, A.; Yariv, A. (2014). "High-coherence semiconductor lasers based on integral high-Q resonators in hybrid Si/III-V platforms". Proceedings of the National Academy of Sciences of the United States of America. 111 (8): 2879–2884. Bibcode:2014PNAS..111.2879S. doi: 10.1073/pnas.1400184111 . PMC   3939879 . PMID   24516134.
  27. L. W. Hollberg, CW dye lasers, in Dye Laser Principles, F. J. Duarte and L. W. Hillman (eds.) (Academic, New York, 1990) Chapter 5.
  28. Spaeth, M. L.; Bortfeld, D. P. (1966). "Stimulated emission from polymethine dyes". Applied Physics Letters. 9 (5). AIP Publishing: 179–181. Bibcode:1966ApPhL...9..179S. doi:10.1063/1.1754699. ISSN   0003-6951.
  29. 1 2 F. J. Duarte,Tunable Laser Optics, 2nd Edition (CRC, New York, 2015).
  30. J. K. Robertson, Introduction to Optics: Geometrical and Physical (Van Nostrand, New York, 1955).
  31. Duarte, F. J. (1992-11-20). "Cavity dispersion equation Δλ ≈ Δθ(∂θ/∂λ)−1: a note on its origin". Applied Optics. 31 (33). The Optical Society: 6979–82. doi:10.1364/ao.31.006979. ISSN   0003-6935. PMID   20802556.
  32. 1 2 Duarte, Francisco J. (1999-10-20). "Multiple-prism grating solid-state dye laser oscillator: optimized architecture". Applied Optics. 38 (30). The Optical Society: 6347–9. Bibcode:1999ApOpt..38.6347D. doi:10.1364/ao.38.006347. ISSN   0003-6935. PMID   18324163.