Optical cavity

Last updated

An optical cavity, resonating cavity or optical resonator is an arrangement of mirrors or other optical elements that forms a cavity resonator for light waves. Optical cavities are a major component of lasers, surrounding the gain medium and providing feedback of the laser light. They are also used in optical parametric oscillators and some interferometers. Light confined in the cavity reflects multiple times, producing modes with certain resonance frequencies. Modes can be decomposed into longitudinal modes that differ only in frequency and transverse modes that have different intensity patterns across the cross-section of the beam. Many types of optical cavity produce standing wave modes.

Contents

A glass nanoparticle is suspended in an optical cavity Glass nanoparticle suspended in an optical cavity.png
A glass nanoparticle is suspended in an optical cavity

Different resonator types are distinguished by the focal lengths of the two mirrors and the distance between them. Flat mirrors are not often used because of the difficulty of aligning them to the needed precision. The geometry (resonator type) must be chosen so that the beam remains stable, i.e. the size of the beam does not continually grow with multiple reflections. Resonator types are also designed to meet other criteria such as minimum beam waist or having no focal point (and therefore intense light at that point) inside the cavity.

Optical cavities are designed to have a large Q factor; [1] a beam will reflect a very large number of times with little attenuation. Therefore, the frequency line width of the beam is very small indeed compared to the frequency of the laser.

Resonator modes

Types of two-mirror optical cavities, with mirrors of various curvatures, showing the radiation pattern inside each cavity. Optical-cavity1.png
Types of two-mirror optical cavities, with mirrors of various curvatures, showing the radiation pattern inside each cavity.

Light confined in a resonator will reflect multiple times from the mirrors, and due to the effects of interference, only certain patterns and frequencies of radiation will be sustained by the resonator, with the others being suppressed by destructive interference. In general, radiation patterns which are reproduced on every round-trip of the light through the resonator are the most stable, and these are the eigenmodes, known as the modes, of the resonator. [2]

Resonator modes can be divided into two types: longitudinal modes, which differ in frequency from each other; and transverse modes, which may differ in both frequency and the intensity pattern of the light. The basic, or fundamental transverse mode of a resonator is a Gaussian beam.

Resonator types


The most common types of optical cavities consist of two facing plane (flat) or spherical mirrors. The simplest of these is the plane-parallel or Fabry–Pérot cavity, consisting of two opposing flat mirrors. [3] [4] [5] [6] [7] [8] [9] While simple, this arrangement is rarely used in large-scale lasers due to the difficulty of alignment; the mirrors must be aligned parallel within a few seconds of arc, or "walkoff" of the intracavity beam will result in it spilling out of the sides of the cavity. However, this problem is much reduced for very short cavities with a small mirror separation distance (L < 1 cm). Plane-parallel resonators are therefore commonly used in microchip and microcavity lasers and semiconductor lasers. In these cases, rather than using separate mirrors, a reflective optical coating may be directly applied to the laser medium itself. The plane-parallel resonator is also the basis of the Fabry–Pérot interferometer.

For a resonator with two mirrors with radii of curvature R1 and R2, there are a number of common cavity configurations. If the two radii are equal to half the cavity length (R1 = R2 = L / 2), a concentric or spherical resonator results. This type of cavity produces a diffraction-limited beam waist in the centre of the cavity, with large beam diameters at the mirrors, filling the whole mirror aperture. Similar to this is the hemispherical cavity, with one plane mirror and one mirror of radius equal to the cavity length.

A common and important design is the confocal resonator, with mirrors of equal radii to the cavity length (R1 = R2 = L). [10] [11] [12] [13] [14] [15] This design produces the smallest possible beam diameter at the cavity mirrors for a given cavity length, and is often used in lasers where the purity of the transverse mode pattern is important.

A concave-convex cavity has one convex mirror with a negative radius of curvature. This design produces no intracavity focus of the beam, and is thus useful in very high-power lasers where the intensity of the intracavity light might be damaging to the intracavity medium if brought to a focus.

Spherical cavity

A transparent dielectric sphere, such as a liquid droplet, also forms an interesting optical cavity. In 1986 Richard K. Chang et al. demonstrated lasing using ethanol microdroplets (20–40 micrometers in radius) doped with rhodamine 6G dye. [16] This type of optical cavity exhibits optical resonances when the size of the sphere or the optical wavelength or the refractive index is varied. The resonance is known as morphology-dependent resonance.

Stability

Stability diagram for a two-mirror cavity. Blue-shaded areas correspond to stable configurations. Laser resonator stability.svg
Stability diagram for a two-mirror cavity. Blue-shaded areas correspond to stable configurations.

Only certain ranges of values for R1, R2, and L produce stable resonators in which periodic refocussing of the intracavity beam is produced. If the cavity is unstable, the beam size will grow without limit, eventually growing larger than the size of the cavity mirrors and being lost. By using methods such as ray transfer matrix analysis, it is possible to calculate a stability criterion: [17]

Values which satisfy the inequality correspond to stable resonators.

The stability can be shown graphically by defining a stability parameter, g for each mirror:

,

and plotting g1 against g2 as shown. Areas bounded by the line g1g2 = 1 and the axes are stable. Cavities at points exactly on the line are marginally stable; small variations in cavity length can cause the resonator to become unstable, and so lasers using these cavities are in practice often operated just inside the stability line.

A simple geometric statement describes the regions of stability: A cavity is stable if the line segments between the mirrors and their centers of curvature overlap, but one does not lie entirely within the other.

In the confocal cavity, if a ray is deviated from its original direction in the middle of the cavity, its displacement after reflecting from one of the mirrors is larger than in any other cavity design. This prevents amplified spontaneous emission and is important for designing high power amplifiers with good beam quality.

Practical resonators

If the optical cavity is not empty (e.g., a laser cavity which contains the gain medium), the value of L used is the physical mirror separation, not the optical path length between the mirrors. Optical elements such as lenses placed in the cavity alter the stability and mode size. In addition, for most gain media, thermal and other inhomogeneities create a variable lensing effect in the medium, which must be considered in the design of the laser resonator.

Practical laser resonators may contain more than two mirrors; three- and four-mirror arrangements are common, producing a "folded cavity". Commonly, a pair of curved mirrors form one or more confocal sections, with the rest of the cavity being quasi-collimated and using plane mirrors. The shape of the laser beam depends on the type of resonator: The beam produced by stable, paraxial resonators can be well modeled by a Gaussian beam. In special cases the beam can be described as a single transverse mode and the spatial properties can be well described by the Gaussian beam, itself. More generally, this beam may be described as a superposition of transverse modes. Accurate description of such a beam involves expansion over some complete, orthogonal set of functions (over two-dimensions) such as Hermite polynomials or the Ince polynomials. Unstable laser resonators on the other hand, have been shown to produce fractal shaped beams. [18]

Some intracavity elements are usually placed at a beam waist between folded sections. Examples include acousto-optic modulators for cavity dumping and vacuum spatial filters for transverse mode control. For some low power lasers, the laser gain medium itself may be positioned at a beam waist. Other elements, such as filters, prisms and diffraction gratings often need large quasi-collimated beams.

These designs allow compensation of the cavity beam's astigmatism, which is produced by Brewster-cut elements in the cavity. A 'Z'-shaped arrangement of the cavity also compensates for coma while the 'delta' or 'X'-shaped cavity does not.

Out of plane resonators lead to rotation of the beam profile and more stability. The heat generated in the gain medium leads to frequency drift of the cavity, therefore the frequency can be actively stabilized by locking it to unpowered cavity. Similarly the pointing stability of a laser may still be improved by spatial filtering by an optical fibre.

Alignment

Alignment of a folded cavity using an autocollimator Metrology-system-paper-4.jpg
Alignment of a folded cavity using an autocollimator

Precise alignment is important when assembling an optical cavity. For best output power and beam quality, optical elements must be aligned such that the path followed by the beam is centered through each element.

Simple cavities are often aligned with an alignment laser—a well-collimated visible laser that can be directed along the axis of the cavity. Observation of the path of the beam and its reflections from various optical elements allows the elements' positions and tilts to be adjusted.

More complex cavities may be aligned using devices such as electronic autocollimators and laser beam profilers.

Optical delay lines

Optical cavities can also be used as multipass optical delay lines, folding a light beam so that a long path-length may be achieved in a small space. A plane-parallel cavity with flat mirrors produces a flat zigzag light path, but as discussed above, these designs are very sensitive to mechanical disturbances and walk-off. When curved mirrors are used in a nearly confocal configuration, the beam travels on a circular zigzag path. The latter is called a Herriott-type delay line. A fixed insertion mirror is placed off-axis near one of the curved mirrors, and a mobile pickup mirror is similarly placed near the other curved mirror. A flat linear stage with one pickup mirror is used in case of flat mirrors and a rotational stage with two mirrors is used for the Herriott-type delay line.

The rotation of the beam inside the cavity alters the polarization state of the beam. To compensate for this, a single pass delay line is also needed, made of either a three or two mirrors in a 3d respective 2d retro-reflection configuration on top of a linear stage. To adjust for beam divergence a second car on the linear stage with two lenses can be used. The two lenses act as a telescope producing a flat phase front of a Gaussian beam on a virtual end mirror.

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 acronym for "light amplification by stimulated emission of radiation". The first laser was built in 1960 by Theodore H. Maiman at Hughes Research Laboratories, based on theoretical work by Charles Hard Townes and Arthur Leonard Schawlow.

<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">Interferometry</span> Measurement method using interference of waves

Interferometry is a technique which uses the interference of superimposed waves to extract information. Interferometry typically uses electromagnetic waves and is an important investigative technique in the fields of astronomy, fiber optics, engineering metrology, optical metrology, oceanography, seismology, spectroscopy, quantum mechanics, nuclear and particle physics, plasma physics, remote sensing, biomolecular interactions, surface profiling, microfluidics, mechanical stress/strain measurement, velocimetry, optometry, and making holograms.

<span class="mw-page-title-main">Fabry–Pérot interferometer</span>

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".

Mode locking is a technique in optics by which a laser can be made to produce pulses of light of extremely short duration, on the order of picoseconds (10−12 s) or femtoseconds (10−15 s). A laser operated in this way is sometimes referred to as a femtosecond laser, for example, in modern refractive surgery. The basis of the technique is to induce a fixed phase relationship between the longitudinal modes of the laser's resonant cavity. Constructive interference between these modes can cause the laser light to be produced as a train of pulses. The laser is then said to be "phase-locked" or "mode-locked".

<span class="mw-page-title-main">Dye laser</span> Equipment using an organic dye to emit coherent light

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. 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.

<span class="mw-page-title-main">Vertical-cavity surface-emitting laser</span>

The vertical-cavity surface-emitting laser, or VCSEL, is a type of semiconductor laser diode with laser beam emission perpendicular from the top surface, contrary to conventional edge-emitting semiconductor lasers which emit from surfaces formed by cleaving the individual chip out of a wafer. VCSELs are used in various laser products, including computer mice, fiber optic communications, laser printers, Face ID, and smartglasses.

<span class="mw-page-title-main">Kerr-lens modelocking</span> Laser mode-locking method

Kerr-lens mode-locking (KLM) is a method of mode-locking lasers via the nonlinear optical Kerr effect. This method allows the generation of pulses of light with a duration as short as a few femtoseconds.

<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">Distributed Bragg reflector</span> Structure used in waveguides

A distributed Bragg reflector (DBR) is a reflector used in waveguides, such as optical fibers. It is a structure formed from multiple layers of alternating materials with varying refractive index, or by periodic variation of some characteristic of a dielectric waveguide, resulting in periodic variation in the effective refractive index in the guide. Each layer boundary causes a partial reflection of an optical wave. For waves whose vacuum wavelength is close to four times the optical thickness of the layers, the many reflections combine with constructive interference, and the layers act as a high-quality reflector. The range of wavelengths that are reflected is called the photonic stopband. Within this range of wavelengths, light is "forbidden" to propagate in the structure.

This is a list of acronyms and other initialisms used in laser physics and laser applications.

<span class="mw-page-title-main">Sound amplification by stimulated emission of radiation</span>

Sound amplification by stimulated emission of radiation (SASER) refers to a device that emits acoustic radiation. It focuses sound waves in a way that they can serve as accurate and high-speed carriers of information in many kinds of applications—similar to uses of laser light.

<span class="mw-page-title-main">Output coupler</span>

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.

A distributed-feedback laser (DFB) is a type of laser diode, quantum-cascade laser or optical-fiber laser where the active region of the device contains a periodically structured element or diffraction grating. The structure builds a one-dimensional interference grating, and the grating provides optical feedback for the laser. This longitudinal diffraction grating has periodic changes in refractive index that cause reflection back into the cavity. The periodic change can be either in the real part of the refractive index or in the imaginary part. The strongest grating operates in the first order, where the periodicity is one-half wave, and the light is reflected backwards. DFB lasers tend to be much more stable than Fabry–Perot or DBR lasers and are used frequently when clean single-mode operation is needed, especially in high-speed fiber-optic telecommunications. Semiconductor DFB lasers in the lowest loss window of optical fibers at about 1.55 μm wavelength, amplified by erbium-doped fiber amplifiers (EDFAs), dominate the long-distance communication market, while DFB lasers in the lowest dispersion window at 1.3 μm are used at shorter distances.

Laser absorption spectrometry (LAS) refers to techniques that use lasers to assess the concentration or amount of a species in gas phase by absorption spectrometry (AS).

The Pound–Drever–Hall (PDH) technique is a widely used and powerful approach for stabilizing the frequency of light emitted by a laser by means of locking to a stable cavity. The PDH technique has a broad range of applications including interferometric gravitational wave detectors, atomic physics, and time measurement standards, many of which also use related techniques such as frequency modulation spectroscopy. Named after R. V. Pound, Ronald Drever, and John L. Hall, the PDH technique was described in 1983 by Drever, Hall and others working at the University of Glasgow and the U. S. National Bureau of Standards. This optical technique has many similarities to an older frequency-modulation technique developed by Pound for microwave cavities.

Anthony E. Siegman was an electrical engineer and educator at Stanford University who investigated and taught about masers and lasers. Known to almost all as Tony Siegman, he was president of the Optical Society of America [now Optica (society)] in 1999 and was awarded the Esther Hoffman Beller Medal in 2009.

Laser linewidth is the spectral linewidth of a laser beam.

Vernier spectroscopy is a type of cavity enhanced laser absorption spectroscopy that is especially sensitive to trace gases. The method uses a frequency comb laser combined with a high finesse optical cavity to produce an absorption spectrum in a highly parallel manner. The method is also capable of detecting trace gases in very low concentration due to the enhancement effect of the optical resonator on the effective optical path length.

<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.

References

  1. Paschotta, Rüdiger. "Q Factor". Encyclopedia of Laser Physics and Technology. RP Photonics.
  2. Lotsch, H.K.V. (1967). "The scalar theory for optical resonators and beam waveguides". Optik. 26: 112–130.
  3. Fox, A.G.; Li, T. (1961). "Resonant modes in a maser interferometer". Bell Syst. Tech. J. 40 (2): 453–488. doi:10.1002/j.1538-7305.1961.tb01625.x.
  4. 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". Optics Express. 24 (15): 16366–16389. Bibcode:2016OExpr..2416366I. doi: 10.1364/OE.24.016366 . PMID   27464090.
  5. Lotsch, H.K.V. (1968). "The Fabry-Perot resonator Part I". Optik. 28: 65–75.
  6. Lotsch, H.K.V. (1969). "The Fabry-Perot resonator Part II". Optik. 28: 328–345.
  7. Lotsch, H.K.V. (1969). "The Fabry-Perot resonator. Part III". Optik. 28: 555–574.
  8. Lotsch, H.K.V. (1969). "The Fabry-Perot resonator. Part IV". Optik. 29: 130–145.
  9. Lotsch, H.K.V. (1969). "The Fabry-Perot resonator Part V". Optik. 29: 622–623.
  10. Boyd, G.D.; Gordon, J.P. (1961). "Confocal multimode resonator for millimeter through optical wavelength masers". Bell Syst. Tech. J. 40 (2): 489–508. doi:10.1002/j.1538-7305.1961.tb01626.x.
  11. Boyd, G.D.; Kogelnik, H. (1962). "Generalized confocal resonator theory". Bell Syst. Tech. J. 41 (4): 1347–1369. doi:10.1002/j.1538-7305.1962.tb03281.x.
  12. Lotsch, H.K.V. (1969). "The confocal resonator system I". Optik. 30: 1–14.
  13. Lotsch, H.K.V. (1969). "The confocal resonator system II". Optik. 30: 181–201.
  14. Lotsch, H.K.V. (1970). "The confocal resonator system III". Optik. 30: 217–233.
  15. Lotsch, H.K.V. (1970). "The confocal resonator system IV". Optik. 30 (6): 563–576.
  16. Qian, Shi-Xiong; Snow, Judith B.; Tzeng, Huey-Ming; Chang, Richard K. (1986-01-31). "Lasing Droplets: Highlighting the Liquid-Air Interface by Laser Emission". Science. 231 (4737): 486–488. Bibcode:1986Sci...231..486Q. doi:10.1126/science.231.4737.486. PMID   17776021. S2CID   5925624.
  17. Yariv, Amnon (1989). Quantum Electronics (3rd ed.). Wiley. p. 142. ISBN   0-4716-0997-8.
  18. Karman, G. P.; et al. (1999). "Laser optics: Fractal modes in unstable resonators". Nature. 402 (6758): 138. Bibcode:1999Natur.402..138K. doi: 10.1038/45960 . S2CID   205046813.
  19. Aharon. "Metrology System for Inter-Alignment of Lasers, Telescopes, and Mechanical Datum".

Further reading