Interferómetro de Fabry-Pérot


En óptica , un interferómetro de Fabry-Pérot ( FPI ) o etalon es una cavidad óptica hecha de dos superficies reflectantes paralelas (es decir, espejos delgados ). Las ondas ópticas pueden atravesar la cavidad óptica solo cuando están en resonancia con ella. Lleva el nombre de Charles Fabry y Alfred Perot , quienes desarrollaron el instrumento en 1899. [1] [2] [3] Etalon proviene del francés étalon , que significa "calibre de medición" o "estándar". [4]

Franjas de interferencia, mostrando una estructura fina , de un etalón de Fabry-Pérot. La fuente es una lámpara de deuterio enfriada .

Los etalones se utilizan ampliamente en telecomunicaciones , láseres y espectroscopia para controlar y medir las longitudes de onda de la luz. Los avances recientes en la técnica de fabricación permiten la creación de interferómetros Fabry-Pérot sintonizables muy precisos. El dispositivo es técnicamente un interferómetro cuando se puede cambiar la distancia entre las dos superficies (y con él la longitud de resonancia), y un etalón cuando la distancia es fija (sin embargo, los dos términos a menudo se usan indistintamente).

Interferómetro de Fabry-Pérot, utilizando un par de planos ópticos parcialmente reflectantes y ligeramente acuñados. El ángulo de la cuña está muy exagerado en esta ilustración; sólo se necesita una fracción de grado para evitar las franjas fantasma. Las imágenes de baja delicadeza frente a las de alta delicadeza corresponden a reflejos de espejo del 4% (vidrio desnudo) y del 95%.

El corazón del interferómetro de Fabry-Pérot es un par de planos ópticos de vidrio parcialmente reflectantes espaciados de micrómetros a centímetros, con las superficies reflectantes enfrentadas entre sí. (Alternativamente, un etalón de Fabry-Pérot usa una sola placa con dos superficies reflectantes paralelas). Los planos en un interferómetro a menudo se hacen en forma de cuña para evitar que las superficies traseras produzcan franjas de interferencia; las superficies traseras a menudo también tienen un revestimiento antirreflectante .

En un sistema típico, la iluminación es proporcionada por una fuente difusa colocada en el plano focal de una lente colimadora . Una lente de enfoque después del par de planos produciría una imagen invertida de la fuente si los planos no estuvieran presentes; toda la luz emitida desde un punto de la fuente se enfoca en un solo punto en el plano de imagen del sistema. En la ilustración adjunta, solo se traza un rayo emitido desde el punto A en la fuente. A medida que el rayo pasa a través de los planos emparejados, se refleja de forma múltiple para producir múltiples rayos transmitidos que son recogidos por la lente de enfoque y llevados al punto A 'en la pantalla. El patrón de interferencia completo toma la apariencia de un conjunto de anillos concéntricos. La nitidez de los anillos depende de la reflectividad de los planos. Si la reflectividad es alta, lo que da como resultado un factor Q alto , la luz monocromática produce un conjunto de anillos estrechos y brillantes sobre un fondo oscuro. Se dice que un interferómetro de Fabry-Pérot con Q alto tiene una gran delicadeza .

Un dispositivo comercial de Fabry-Perot
  • Las redes de telecomunicaciones que emplean multiplexación por división de longitud de onda tienen multiplexores add-drop con bancos de etalones de diamante o sílice fundida sintonizada en miniatura . Se trata de pequeños cubos iridiscentes de unos 2 mm de lado, montados en pequeños bastidores de alta precisión. Los materiales se eligen para mantener distancias estables de espejo a espejo y para mantener frecuencias estables incluso cuando la temperatura varía. Se prefiere el diamante porque tiene una mayor conducción de calor y todavía tiene un bajo coeficiente de expansión. En 2005, algunas empresas de equipos de telecomunicaciones comenzaron a utilizar etalones sólidos que a su vez son fibras ópticas. Esto elimina la mayoría de las dificultades de montaje, alineación y enfriamiento.
  • Los filtros dicroicos se fabrican depositando una serie de capas etalónicas sobre una superficie óptica mediante deposición de vapor . Estos filtros ópticos suelen tener bandas reflectantes y de paso más exactas que los filtros absorbentes. Cuando se diseñan correctamente, funcionan a menor temperatura que los filtros absorbentes porque pueden reflejar longitudes de onda no deseadas. Los filtros dicroicos se utilizan ampliamente en equipos ópticos como fuentes de luz, cámaras, equipos astronómicos y sistemas láser.
  • Los medidores de onda ópticos y algunos analizadores de espectro óptico utilizan interferómetros Fabry-Pérot con diferentes rangos espectrales libres para determinar la longitud de onda de la luz con gran precisión.
  • Los resonadores láser a menudo se describen como resonadores Fabry-Pérot, aunque para muchos tipos de láser la reflectividad de un espejo es cercana al 100%, lo que lo hace más similar a un interferómetro Gires-Tournois . Los láseres de diodos semiconductores a veces utilizan una verdadera geometría Fabry-Pérot, debido a la dificultad de recubrir las caras de los extremos del chip. Los láseres de cascada cuántica a menudo emplean cavidades de Fabry-Pérot para sostener el láser sin la necesidad de ningún recubrimiento de facetas, debido a la alta ganancia de la región activa. [5]
  • Los etalones a menudo se colocan dentro del resonador láser cuando se construyen láseres monomodo. Sin un etalón, un láser generalmente producirá luz en un rango de longitud de onda correspondiente a varios modos de cavidad , que son similares a los modos Fabry-Pérot. Insertar un etalón en la cavidad del láser, con una delicadeza bien elegida y un rango espectral libre, puede suprimir todos los modos de la cavidad excepto uno, cambiando así el funcionamiento del láser de multimodo a monomodo.
  • Los etalones de Fabry-Pérot se pueden utilizar para prolongar la duración de la interacción en la espectrometría de absorción láser , en particular en las técnicas de reducción de la cavidad .
  • Se puede usar un etalón de Fabry-Pérot para hacer un espectrómetro capaz de observar el efecto Zeeman , donde las líneas espectrales están demasiado juntas para distinguirlas con un espectrómetro normal.
  • En astronomía, un etalón se utiliza para seleccionar una única transición atómica para la obtención de imágenes. La más común es la línea H-alfa del sol . La línea Ca-K del sol también se representa comúnmente usando etalones.
  • El sensor de metano para Marte (MSM) a bordo del Mangalyaan de la India es un ejemplo de un instrumento Fabry-Perot. Fue el primer instrumento de Fabry Perot en el espacio cuando se lanzó Mangalyaan. [6] Como no distinguía la radiación absorbida por el metano de la radiación absorbida por el dióxido de carbono y otros gases, más tarde se denominó mapeador de albedo. [7]
  • En la detección de ondas gravitacionales , se utiliza una cavidad de Fabry-Pérot para almacenar fotones durante casi un milisegundo mientras rebotan hacia arriba y hacia abajo entre los espejos. Esto aumenta el tiempo que una onda gravitacional puede interactuar con la luz, lo que da como resultado una mejor sensibilidad a bajas frecuencias. Este principio es utilizado por detectores como LIGO y Virgo , que consisten en un interferómetro de Michelson con una cavidad de Fabry-Pérot con una longitud de varios kilómetros en ambos brazos. Las cavidades más pequeñas, generalmente llamadas limpiadores de modo , se utilizan para el filtrado espacial y la estabilización de frecuencia del láser principal.

Pérdidas de resonador, luz desacoplada, frecuencias de resonancia y formas de líneas espectrales

La respuesta espectral de un resonador de Fabry-Pérot se basa en la interferencia entre la luz lanzada en él y la luz que circula en el resonador. La interferencia constructiva ocurre si los dos haces están en fase , lo que conduce a una mejora resonante de la luz dentro del resonador. Si los dos haces están desfasados, solo una pequeña parte de la luz lanzada se almacena dentro del resonador. La luz almacenada, transmitida y reflejada se modifica espectralmente en comparación con la luz incidente.

Suponga un resonador Fabry-Pérot de dos espejos de longitud geométrica , relleno homogéneamente con un medio de índice de refracción . La luz se lanza al resonador con una incidencia normal. El tiempo de ida y vuelta de luz viajando en el resonador con velocidad , dónde es la velocidad de la luz en el vacío y el rango espectral libre son dadas por

Las reflectividades del campo eléctrico y de la intensidad. y , respectivamente, en el espejo están

Si no hay otras pérdidas de resonador, la caída de la intensidad de la luz por viaje de ida y vuelta se cuantifica mediante la constante de velocidad de caída de desacoplamiento

y el tiempo de desintegración de fotones del resonador viene dado por [8]

Con cuantificar el cambio de fase de un solo paso que exhibe la luz cuando se propaga de un espejo al otro, el cambio de fase de ida y vuelta en frecuencia se acumula en [8]

Las resonancias ocurren a frecuencias en las que la luz exhibe interferencia constructiva después de un viaje de ida y vuelta. Cada modo de resonador con su índice de modo, dónde es un número entero en el intervalo [, ..., −1, 0, 1, ..., ], está asociado con una frecuencia de resonancia y número de onda ,

Dos modos con valores opuestos y de índice modal y número de onda, respectivamente, que representan físicamente direcciones de propagación opuestas, ocurren en el mismo valor absoluto de la frecuencia. [9]

El campo eléctrico en descomposición a frecuencia. está representado por una oscilación armónica amortiguada con una amplitud inicial de y una constante de tiempo de decaimiento de . En notación fasorial, se puede expresar como [8]

La transformación de Fourier del campo eléctrico en el tiempo proporciona el campo eléctrico por intervalo de frecuencia unitario,

Cada modo tiene una forma de línea espectral normalizada por intervalo de frecuencia unitario dado por

cuya integral de frecuencia es la unidad. Presentamos el ancho de línea de ancho completo a la mitad del máximo (FWHM) de la forma de la línea espectral de Lorentz, obtenemos

expresado en términos de ancho de línea de medio ancho a medio máximo (HWHM) o el ancho de línea FWHM . Calibrado a una altura máxima de la unidad, obtenemos las líneas de Lorentzian:

Al repetir la transformación de Fourier anterior para todos los modos con índice de modo en el resonador, se obtiene el espectro de modo completo del resonador.

Dado que el ancho de línea y el rango espectral libre son independientes de la frecuencia, mientras que en el espacio de longitud de onda el ancho de línea no se puede definir correctamente y el rango espectral libre depende de la longitud de onda, y dado que las frecuencias de resonancia escala proporcional a la frecuencia, la respuesta espectral de un resonador Fabry-Pérot se analiza y muestra de forma natural en el espacio de frecuencia.

Distribución genérica de Airy: el factor de mejora de la resonancia interna

Caption
Campos eléctricos en un resonador Fabry-Pérot. [8] Las reflectividades del espejo de campo eléctrico son y . Se indican los campos eléctricos característicos producidos por un campo eléctrico. incidente en el espejo 1: inicialmente reflejado en el espejo 1, lanzado a través del espejo 1, y circulando dentro del resonador en dirección de propagación hacia adelante y hacia atrás, respectivamente, propagarse dentro del resonador después de un viaje de ida y vuelta, transmitido a través del espejo 2, transmitido a través del espejo 1, y el campo total propagándose hacia atrás. La interferencia ocurre en los lados izquierdo y derecho del espejo 1 entre y , Resultando en y entre y , Resultando en , respectivamente.

La respuesta del resonador de Fabry-Pérot a un campo eléctrico que incide en el espejo 1 se describe mediante varias distribuciones de Airy (nombradas en honor al matemático y astrónomo George Biddell Airy ) que cuantifican la intensidad de la luz en la dirección de propagación hacia adelante o hacia atrás en diferentes posiciones dentro o fuera el resonador con respecto a la intensidad de la luz lanzada o incidente. La respuesta del resonador de Fabry-Pérot se obtiene más fácilmente mediante el uso del enfoque de campo circulante. [10] Este enfoque asume un estado estable y relaciona los diversos campos eléctricos entre sí (ver figura "Campos eléctricos en un resonador Fabry-Pérot").

El campo can be related to the field that is launched into the resonator by

The generic Airy distribution, which considers solely the physical processes exhibited by light inside the resonator, then derives as the intensity circulating in the resonator relative to the intensity launched,[8]

represents the spectrally dependent internal resonance enhancement which the resonator provides to the light launched into it (see figure "Resonance enhancement in a Fabry-Pérot resonator"). At the resonance frequencies , where equals zero, the internal resonance enhancement factor is

Other Airy distributions

Caption
Resonance enhancement in a Fabry-Pérot resonator. [8] (top) Spectrally dependent internal resonance enhancement, equaling the generic Airy distribution . Light launched into the resonator is resonantly enhanced by this factor. For the curve with , the peak value is at , outside the scale of the ordinate. (bottom) Spectrally dependent external resonance enhancement, equaling the Airy distribution . Light incident upon the resonator is resonantly enhanced by this factor.

Once the internal resonance enhancement, the generic Airy distribution, is established, all other Airy distributions can be deduced by simple scaling factors.[8] Since the intensity launched into the resonator equals the transmitted fraction of the intensity incident upon mirror 1,

and the intensities transmitted through mirror 2, reflected at mirror 2, and transmitted through mirror 1 are the transmitted and reflected/transmitted fractions of the intensity circulating inside the resonator,

respectively, the other Airy distributions with respect to launched intensity and with respect to incident intensity are[8]

The index "emit" denotes Airy distributions that consider the sum of intensities emitted on both sides of the resonator.

The back-transmitted intensity cannot be measured, because also the initially back-reflected light adds to the backward-propagating signal. The measurable case of the intensity resulting from the interference of both backward-propagating electric fields results in the Airy distribution[8]

It can be easily shown that in a Fabry-Pérot resonator, despite the occurrence of constructive and destructive interference, energy is conserved at all frequencies:

The external resonance enhancement factor (see figure "Resonance enhancement in a Fabry-Pérot resonator") is[8]

At the resonance frequencies , where equals zero, the external resonance enhancement factor is

Caption
Airy distribution (solid lines), corresponding to light transmitted through a Fabry-Pérot resonator, calculated for different values of the reflectivities , and comparison with a single Lorentzian line (dashed lines) calculated for the same . [8] At half maximum (black line), with decreasing reflectivities the FWHM linewidth of the Airy distribution broadens compared to the FWHM linewidth of its corresponding Lorentzian line: results in , respectively.

Usually light is transmitted through a Fabry-Pérot resonator. Therefore, an often applied Airy distribution is[8]

It describes the fraction of the intensity of a light source incident upon mirror 1 that is transmitted through mirror 2 (see figure "Airy distribution "). Its peak value at the resonance frequencies is

For the peak value equals unity, i.e., all light incident upon the resonator is transmitted; consequently, no light is reflected, , as a result of destructive interference between the fields and .

has been derived in the circulating-field approach[10] by considering an additional phase shift of during each transmission through a mirror,

resulting in

Alternatively, can be obtained via the round-trip-decay approach[11] by tracing the infinite number of round trips that the incident electric field exhibits after entering the resonator and accumulating the electric field transmitted in all round trips. The field transmitted after the first propagation and the smaller and smaller fields transmitted after each consecutive propagation through the resonator are

respectively. Exploiting

results in the same as above, therefore the same Airy distribution derives. However, this approach is physically misleading, because it assumes that interference takes place between the outcoupled beams after mirror 2, outside the resonator, rather than the launched and circulating beams after mirror 1, inside the resonator. Since it is interference that modifies the spectral contents, the spectral intensity distribution inside the resonator would be the same as the incident spectral intensity distribution, and no resonance enhancement would occur inside the resonator.

Airy distribution as a sum of mode profiles

Physically, the Airy distribution is the sum of mode profiles of the longitudinal resonator modes.[8] Starting from the electric field circulating inside the resonator, one considers the exponential decay in time of this field through both mirrors of the resonator, Fourier transforms it to frequency space to obtain the normalized spectral line shapes , divides it by the round-trip time to account for how the total circulating electric-field intensity is longitudinally distributed in the resonator and coupled out per unit time, resulting in the emitted mode profiles,

and then sums over the emitted mode profiles of all longitudinal modes[8]

thus equaling the Airy distribution .

The same simple scaling factors that provide the relations between the individual Airy distributions also provide the relations among and the other mode profiles:[8]

Characterizing the Fabry-Pérot resonator: Lorentzian linewidth and finesse

The Taylor criterion of spectral resolution proposes that two spectral lines can be resolved if the individual lines cross at half intensity. When launching light into the Fabry-Pérot resonator, by measuring the Airy distribution, one can derive the total loss of the Fabry-Pérot resonator via recalculating the Lorentzian linewidth , displayed (blue line) relative to the free spectral range in the figure "Lorentzian linewidth and finesse versus Airy linewidth and finesse of a Fabry-Pérot resonator".

Caption
Lorentzian linewidth and finesse versus Airy linewidth and finesse of a Fabry-Pérot resonator. [8] [Left] Relative Lorentzian linewidth (blue curve), relative Airy linewidth (green curve), and its approximation (red curve). [Right] Lorentzian finesse (blue curve), Airy finesse (green curve), and its approximation (red curve) as a function of reflectivity value . The exact solutions of the Airy linewidth and finesse (green lines) correctly break down at , equivalent to , whereas their approximations (red lines) incorrectly do not break down. Insets: Region .
Caption
The physical meaning of the Lorentzian finesse of a Fabry-Pérot resonator. [8] Displayed is the situation for , at which and , i.e., two adjacent Lorentzian lines (dashed colored lines, only 5 lines are shown for clarity for each resonance frequency, ) cross at half maximum (solid black line) and the Taylor criterion for spectrally resolving two peaks in the resulting Airy distribution (solid purple line, the sum of 5 lines which has been normalized to the peak intensity of itself) is reached.

The underlying Lorentzian lines can be resolved as long as the Taylor criterion is obeyed (see figure "The physical meaning of the Lorentzian finesse"). Consequently, one can define the Lorentzian finesse of a Fabry-Pérot resonator:[8]

It is displayed as the blue line in the figure "The physical meaning of the Lorentzian finesse". The Lorentzian finesse has a fundamental physical meaning: it describes how well the Lorentzian lines underlying the Airy distribution can be resolved when measuring the Airy distribution. At the point where

equivalent to , the Taylor criterion for the spectral resolution of a single Airy distribution is reached. Under this point, , two spectral lines cannot be distinguished. For equal mirror reflectivities, this point occurs when . Therefore, the linewidth of the Lorentzian lines underlying the Airy distribution of a Fabry-Pérot resonator can be resolved by measuring the Airy distribution, hence its resonator losses can be spectroscopically determined, until this point.

Scanning the Fabry-Pérot resonator: Airy linewidth and finesse

Caption
The physical meaning of the Airy finesse of a Fabry-Pérot resonator. [8] When scanning the Fabry-Pérot length (or the angle of incident light), Airy distributions (colored solid lines) are created by signals at individual frequencies. The experimental result of the measurement is the sum of the individual Airy distributions (black dashed line). If the signals occur at frequencies , where is an integer starting at , the Airy distributions at adjacent frequencies are separated from each other by the linewidth , thereby fulfilling the Taylor criterion for the spectroscopic resolution of two adjacent peaks. The maximum number of signals that can be resolved is . Since in this specific example the reflectivities have been chosen such that is an integer, the signal for at the frequency coincides with the signal for at . In this example, a maximum of peaks can be resolved when applying the Taylor criterion.
Caption
Example of a Fabry-Pérot resonator with (top) frequency-dependent mirror reflectivity and (bottom) the resulting distorted mode profiles of the modes with indices , the sum of 6 million mode profiles (pink dots, displayed for a few frequencies only), and the Airy distribution . [8] The vertical dashed lines denote the maximum of the reflectivity curve (black) and the resonance frequencies of the individual modes (colored).

When the Fabry-Pérot resonator is used as a scanning interferometer, i.e., at varying resonator length (or angle of incidence), one can spectroscopically distinguish spectral lines at different frequencies within one free spectral range. Several Airy distributions , each one created by an individual spectral line, must be resolved. Therefore, the Airy distribution becomes the underlying fundamental function and the measurement delivers a sum of Airy distributions. The parameters that properly quantify this situation are the Airy linewidth and the Airy finesse . The FWHM linewidth of the Airy distribution is[8]

The Airy linewidth is displayed as the green curve in the figure "Lorentzian linewidth and finesse versus Airy linewidth and finesse of a Fabry-Pérot resonator".

The concept of defining the linewidth of the Airy peaks as FWHM breaks down at (solid red line in the figure "Airy distribution "), because at this point the Airy linewidth instantaneously jumps to an infinite value for function. For lower reflectivity values of , the FWHM linewidth of the Airy peaks is undefined. The limiting case occurs at

For equal mirror reflectivities, this point is reached when (solid red line in the figure "Airy distribution ").

The finesse of the Airy distribution of a Fabry-Pérot resonator, which is displayed as the green curve in the figure "Lorentzian linewidth and finesse versus Airy linewidth and finesse of a Fabry-Pérot resonator" in direct comparison with the Lorentzian finesse , is defined as[8]

When scanning the length of the Fabry-Pérot resonator (or the angle of incident light), the Airy finesse quantifies the maximum number of Airy distributions created by light at individual frequencies within the free spectral range of the Fabry-Pérot resonator, whose adjacent peaks can be unambiguously distinguished spectroscopically, i.e., they do not overlap at their FWHM (see figure "The physical meaning of the Airy finesse"). This definition of the Airy finesse is consistent with the Taylor criterion of the resolution of a spectrometer. Since the concept of the FWHM linewidth breaks down at , consequently the Airy finesse is defined only until , see the figure "Lorentzian linewidth and finesse versus Airy linewidth and finesse of a Fabry-Pérot resonator".

Often the unnecessary approximation is made when deriving from the Airy linewidth . In contrast to the exact solution above, it leads to

This approximation of the Airy linewidth, displayed as the red curve in the figure "Lorentzian linewidth and finesse versus Airy linewidth and finesse of a Fabry-Pérot resonator", deviates from the correct curve at low reflectivities and incorrectly does not break down when . This approximation is then typically also used to calculate the Airy finesse.

Frequency-dependent mirror reflectivities

The more general case of a Fabry-Pérot resonator with frequency-dependent mirror reflectivities can be treated with the same equations as above, except that the photon decay time and linewidth now become local functions of frequency. Whereas the photon decay time is still a well-defined quantity, the linewidth loses its meaning, because it resembles a spectral bandwidth, whose value now changes within that very bandwidth. Also in this case each Airy distribution is the sum of all underlying mode profiles which can be strongly distorted.[8] An example of the Airy distribution and a few of the underlying mode profiles is given in the figure "Example of a Fabry-Pérot resonator with frequency-dependent mirror reflectivity".

Fabry-Pérot resonator with intrinsic optical losses

Intrinsic propagation losses inside the resonator can be quantified by an intensity-loss coefficient per unit length or, equivalently, by the intrinsic round-trip loss such that[12]

The additional loss shortens the photon-decay time of the resonator:[12]

where is the light speed in cavity. The generic Airy distribution or internal resonance enhancement factor is then derived as above by including the propagation losses via the amplitude-loss coefficient :[12]

The other Airy distributions can then be derived as above by additionally taking into account the propagation losses. Particularly, the transfer function with loss becomes[12]

Description of the Fabry-Perot resonator in wavelength space

A Fabry–Pérot etalon. Light enters the etalon and undergoes multiple internal reflections.
The transmission of an etalon as a function of wavelength. A high-finesse etalon (red line) shows sharper peaks and lower transmission minima than a low-finesse etalon (blue).
Finesse as a function of reflectivity. Very high finesse factors require highly reflective mirrors.
Fabry Perot Diagram1.svg
"> Play media
Transient analysis of a silicon ( n = 3.4) Fabry–Pérot etalon at normal incidence. The upper animation is for etalon thickness chosen to give maximum transmission while the lower animation is for thickness chosen to give minimum transmission.
"> Play media
False color transient for a high refractive index, dielectric slab in air. The thickness/frequencies have been selected such that red (top) and blue (bottom) experience maximum transmission, whereas the green (middle) experiences minimum transmission.

The varying transmission function of an etalon is caused by interference between the multiple reflections of light between the two reflecting surfaces. Constructive interference occurs if the transmitted beams are in phase, and this corresponds to a high-transmission peak of the etalon. If the transmitted beams are out-of-phase, destructive interference occurs and this corresponds to a transmission minimum. Whether the multiply reflected beams are in phase or not depends on the wavelength (λ) of the light (in vacuum), the angle the light travels through the etalon (θ), the thickness of the etalon () and the refractive index of the material between the reflecting surfaces (n).

The phase difference between each successive transmitted pair (i.e. T2 and T1 in the diagram) is given by[13]

If both surfaces have a reflectance R, the transmittance function of the etalon is given by

where

is the coefficient of finesse.

Maximum transmission () occurs when the optical path length difference () between each transmitted beam is an integer multiple of the wavelength. In the absence of absorption, the reflectance of the etalon Re is the complement of the transmittance, such that . The maximum reflectivity is given by

and this occurs when the path-length difference is equal to half an odd multiple of the wavelength.

The wavelength separation between adjacent transmission peaks is called the free spectral range (FSR) of the etalon, Δλ, and is given by:

where λ0 is the central wavelength of the nearest transmission peak and is the group refractive index.[14] The FSR is related to the full-width half-maximum, δλ, of any one transmission band by a quantity known as the finesse:

This is commonly approximated (for R > 0.5) by

If the two mirrors are not equal, the finesse becomes

Etalons with high finesse show sharper transmission peaks with lower minimum transmission coefficients. In the oblique incidence case, the finesse will depend on the polarization state of the beam, since the value of R, given by the Fresnel equations, is generally different for p and s polarizations.

Two beams are shown in the diagram at the right, one of which (T0) is transmitted through the etalon, and the other of which (T1) is reflected twice before being transmitted. At each reflection, the amplitude is reduced by , while at each transmission through an interface the amplitude is reduced by . Assuming no absorption, conservation of energy requires T + R = 1. In the derivation below, n is the index of refraction inside the etalon, and n0 is that outside the etalon. It is presumed that n > n0. The incident amplitude at point a is taken to be one, and phasors are used to represent the amplitude of the radiation. The transmitted amplitude at point b will then be

where is the wavenumber inside the etalon, and λ is the vacuum wavelength. At point c the transmitted amplitude will be

The total amplitude of both beams will be the sum of the amplitudes of the two beams measured along a line perpendicular to the direction of the beam. The amplitude t0 at point b can therefore be added to t'1 retarded in phase by an amount , where is the wavenumber outside of the etalon. Thus

where ℓ0 is

The phase difference between the two beams is

The relationship between θ and θ0 is given by Snell's law:

so that the phase difference may be written as

To within a constant multiplicative phase factor, the amplitude of the mth transmitted beam can be written as

The total transmitted amplitude is the sum of all individual beams' amplitudes:

The series is a geometric series, whose sum can be expressed analytically. The amplitude can be rewritten as

The intensity of the beam will be just t times its complex conjugate. Since the incident beam was assumed to have an intensity of one, this will also give the transmission function:

For an asymmetrical cavity, that is, one with two different mirrors, the general form of the transmission function is

A Fabry–Pérot interferometer differs from a Fabry–Pérot etalon in the fact that the distance between the plates can be tuned in order to change the wavelengths at which transmission peaks occur in the interferometer. Due to the angle dependence of the transmission, the peaks can also be shifted by rotating the etalon with respect to the beam.

Another expression for the transmission function was already derived in the description in frequency space as the infinite sum of all longitudinal mode profiles. Defining the above expression may be written as

The second term is proportional to a wrapped Lorentzian distribution so that the transmission function may be written as a series of Lorentzian functions:

where

  • Lummer–Gehrcke interferometer
  • Gires–Tournois etalon
  • Atomic line filter
  • ARROW waveguide
  • Distributed Bragg reflector
  • Fiber Bragg grating
  • Optical microcavity
  • Thin-film interference
  • Laser linewidth

  1. ^ Perot frequently spelled his name with an accent—Pérot—in scientific publications, and so the name of the interferometer is commonly written with the accent. Métivier, Françoise (September–October 2006). "Jean-Baptiste Alfred Perot" (PDF). Photoniques (in French) (25). Archived from the original (PDF) on 2007-11-10. Retrieved 2007-10-02. Page 2: "Pérot ou Perot?"
  2. ^ Fabry, C; Perot, A (1899). "Theorie et applications d'une nouvelle methode de spectroscopie interferentielle". Ann. Chim. Phys. 16 (7).
  3. ^ Perot, A; Fabry, C (1899). "On the Application of Interference Phenomena to the Solution of Various Problems of Spectroscopy and Metrology". Astrophysical Journal. 9: 87. Bibcode:1899ApJ.....9...87P. doi:10.1086/140557.
  4. ^ Oxford English Dictionary
  5. ^ Williams, Benjamin S. (2007). "Terahertz quantum-cascade lasers" (PDF). Nature Photonics. 1 (9): 517–525. Bibcode:2007NaPho...1..517W. doi:10.1038/nphoton.2007.166. hdl:1721.1/17012. ISSN 1749-4885. S2CID 29073195.
  6. ^ Mukunth, Vasudevan (2016-12-15). "ISRO Mars Orbiter Mission's Methane Instrument Has a Glitch". The Wire. Retrieved 2019-12-21.
  7. ^ Klotz, Irene (2016-12-07). "India's Mars Orbiter Mission Has a Methane Problem". Seeker.com. Retrieved 2019-12-21.
  8. ^ a b c d e f g h i j k l m n o p q r s t u v w 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.
  9. ^ Pollnau, M. (2018). "Counter-propagating modes in a Fabry-Pérot-type resonator". Optics Letters. 43 (20): 5033–5036. Bibcode:2018OptL...43.5033P. doi:10.1364/OL.43.005033. PMID 30320811.
  10. ^ a b A. E. Siegman, "Lasers", University Science Books, Mill Valley, California, 1986, ch. 11.3, pp. 413-428.
  11. ^ O. Svelto, "Principles of Lasers", 5th ed., Springer, New York, 2010, ch. 4.5.1, pp. 142-146.
  12. ^ a b c d 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. doi:10.1016/j.pquantelec.2020.100255.
  13. ^ Lipson, S. G.; Lipson, H.; Tannhauser, D. S. (1995). Optical Physics (3rd ed.). London: Cambridge U. P. pp. 248. ISBN 0-521-06926-2.
  14. ^ Coldren, L. A.; Corzine, S. W.; Mašanović, M. L. (2012). Diode Lasers and Photonic Integrated Circuits (2nd ed.). Hoboken, New Jersey: Wiley. p. 58. ISBN 978-0-470-48412-8.

  • Hernandez, G. (1986). Fabry–Pérot Interferometers. Cambridge: Cambridge University Press. ISBN 0-521-32238-3.

  • Advanced Design of Etalons- by Precision Photonics Corporation