Sie sind auf Seite 1von 10

Ray transfer matrix analysis (also known as ABCD matrix analysis) is a type of ray tracing technique

used in the design of some optical systems, particularly lasers. It involves the construction of a ray transfer matrix which describes the optical system; tracing of a light path through the system can then be performed by multiplying this matrix with a vector representing the light ray. The same analysis is also used in accelerator physics to track particles trough the magnet installations of a particle accelerator, see Beam optics. The technique that is described below uses the paraxial approximation of ray optics, which means that all rays are assumed to be at a small angle () and a small distance (x) relative to the optical axis of the system.[1] Definition of the ray transfer matrix: The ray tracing technique is based on two reference planes, called the input and output planes, each perpendicular to the optical axis of the system. Without loss of generality, we will define the optical axis so that it coincides with the z-axis of a fixed coordinate system. A light ray enters the system when the ray crosses the input plane at a distance x1 from the optical axis while traveling in a direction that makes an angle 1 with the optical axis. Some distance further along, the ray crosses the output plane, this time at a distance x2 from the optical axis and making an angle 2. n1 and n2 are the indices of refraction of the medium in the input and output plane, respectively. These quantities are related by the expression

where

and

This relates the ray vectors at the input and output planes by the ray transfer matrix (RTM) M, which represents the optical system between the two reference planes. A thermodynamics argument based on the blackbody radiation can be used to show that the determinant of a RTM is the ratio of the indices of refraction:

As a result, if the input and output planes are located within the same medium, or within two different media which happen to have identical indices of refraction, then the determinant of M is simply equal to 1. A similar technique can be used to analyze electrical circuits. See Two-port networks. Some examples

For example, if there is free space between the two planes, the ray transfer matrix is given by:

, where d is the separation distance (measured along the optical axis) between the two reference planes. The ray transfer equation thus becomes:

, and this relates the parameters of the two rays as:

Another simple example is that of a thin lens. Its RTM is given by:

, where f is the focal length of the lens. To describe combinations of optical components, ray transfer matrices may be multiplied together to obtain an overall RTM for the compound optical system. For the example of free space of length d followed by a lens of focal length f:

. Note that, since the multiplication of matrices is non-commutative, this is not the same RTM as that for a lens followed by free space:

. Thus the matrices must be ordered appropriately, with the last matrix premultiplying the second last, and so on until the first matrix is premultiplied by the second. Other matrices can be constructed to represent interfaces with media of different refractive indices, reflection from mirrors, etc. Table of ray transfer matrices for simple optical components Element Propagation in free space or in a medium of constant refractive index Refraction at a flat interface n2 = final refractive index. R = radius of curvature, R > 0 for convex (centre of curvature after interface) Refraction at a curved interface n1 = initial refractive index n2 = final refractive index. Reflection from a flat mirror Matrix d = distance n1 = initial refractive index Remarks

Reflection from a curved mirror

R = radius of curvature, R > 0 for convex

f = focal length of lens where f > 0 for convex/positive (converging) lens. Thin lens Only valid if the focal length is much greater than the thickness of the lens. n1 = refractive index outside of the lens. n2 = refractive index of the lens itself (inside the lens). R1 = Radius of curvature of First surface. R2 = Radius of curvature of Second surface. t = thickness of lens (not taking into account thickness of curved parts. If total thickness is desired, total thickness would equal t + both curved thickness parts. k = (cos /cos ) is the beam expansion factor, where is the angle of incidence, is the angle of refraction, d = prism path length, n = refractive index of the prism material. This matrix applies for orthogonal beam exit.

Thick lens

Single right angle prism

Resonator stability: RTM analysis is particularly useful when modeling the behaviour of light in optical resonators, such as those used in lasers. At its simplest, an optical resonator consists of two identical facing mirrors of 100% reflectivity and radius of curvature R, separated by some distance d. For the purposes of ray tracing, this is equivalent to a series of identical thin lenses of focal length f=R/2, each separated from the next by length d. This construction is known as a lens equivalent duct or lens equivalent waveguide. The RTM of each section of the waveguide is, as above,

. RTM analysis can now be used to determine the stability of the waveguide (and equivalently, the resonator). That is, it can be determined under what conditions light travelling down the waveguide will be periodically refocussed and stay within the waveguide. To do so, we can find all ray vectors where the output of each section of the waveguide is equal to the input vector multiplied by some real or complex constant :

This gives:

which is an eigen value equation: Simplifying, we have which leads to the characteristic equation

where I is the 2x2 identity matrix.

where is the trace of the RTM, and

is the determinant of the RTM. Simplifying, we have

where

is the stability parameter. The eigen values are the solutions of the characteristic equation.

From the quadratic formula, we find After N passes through the system, we have:

. If the waveguide is stable, N must not grow without limit. This observation implies that cannot take on purely real values, but must have a non-zero imaginary part. Thus, As a result, or ,

Solving the eigenvalue equation gives us a periodic solution of the form: Or ,

where The technique may be generalised for more complex resonators by constructing a suitable matrix M for the cavity from the matrices of the components present. Ray transfer matrices for Gaussian beams The matrix formalism is also useful to describe Gaussian beams. If we have a Gaussian beam of wavelength curvature R, beam spot size w and refractive index n, it is possible to define a complex beam parameter q by: . This beam can be propagated through an optical system with a given ray transfer matrix by using the equation: , where k is a normalisation constant chosen to keep the second component of the ray vector equal to 1. Using matrix multiplication, this equation expands as and , radius of

Dividing the first equation by the second eliminates the normalisation constant:

It is often convenient to express this last equation in reciprocal form:

Transfer-matrix method (optics): For the transfer-matrix method in geometric optics, see Ray
transfer matrix analysis. For the transfer-matrix method in statistical physics, see Transfer-matrix method.

The transfer-matrix method is a method used in optics and acoustics to analyze the propagation of electromagnetic or acoustic waves through a stratified (layered) medium.[1] This is for example relevant for the design of anti-reflective coatings and dielectric mirrors. The reflection of light from a single interface between two media is described by the Fresnel equations. However, when there are multiple interfaces, such as in the figure, the reflections themselves are also partially transmitted and then partially reflected. Depending on the exact path length, these reflections can interfere destructively or constructively. The overall reflection of a layer structure is the sum of an infinite number of reflections, which is cumbersome to calculate. The transfer-matrix method is based on the fact that, according to Maxwell's equations, there are simple continuity conditions for the electric field across boundaries from one medium to the next. If the field is known at the beginning of a layer, the field at the end of the layer can be derived from a simple matrix operation. A stack of layers can then be represented as a system matrix, which is the product of the individual layer matrices. The final step of the method involves converting the system matrix back into reflection and transmission coefficients. Formalism for electromagnetic waves: Below is described how the transfer matrix is applied to electromagnetic waves (for example light) of a given frequency propagating through a stack of layers at normal incidence. It can be generalized to deal with incidence at an angle, absorbing media, and media with magnetic properties. We assume that the stack layers are normal to the axis and that the field within one layer can be represented as the superposition of a left- and right-traveling wave with wave number , . Because it follows from Maxwell's equation that and to represent the field as the vector , where . Since there are two equations relating and to and , these two representations are equivalent. In the new representation, propagation over a distance into the positive direction is described by the matrix must be continuous across a boundary, it is convenient

And

Such a matrix can represent propagation through a layer if layer: For a system with layers, each layer system transfer matrix is then

is the wave number in the medium and , where

the thickness of the values. The

has a transfer matrix

increases towards higher

Typically, one would like to know the reflectance and transmittance of the layer structure. If the layer stack starts at then for negative , the field is described as , where is the amplitude of the incoming wave, the wave number in the left medium, and is the amplitude (not intensity!) reflectance coefficient of the layer structure. On the other side of the layer structure, the field consists of a rightpropagating transmitted field , where is the amplitude transmittance and , then we can solve is the wave number in the rightmost medium. If and

in terms of the matrix elements

of the system matrix

and obtain

and

. The transmittance and reflectance (i.e., the fractions of the incident intensity transmitted and reflected by the layer) are often of more practical use and are given by and , respectively. Example As an illustration, consider a single layer of glass with a refractive index n and thickness d suspended in air at a wave number k (in air). In glass, the wave number is . The transfer matrix is

The amplitude reflection coefficient can be simplified to This configuration effectively describes a FabryProt interferometer or etalon: for vanishes.

.
, the reflection

Acoustic waves: It is possible to apply the transfer-matrix method to sound waves. Instead of the electric field E and its derivative F, the displacement u and the stress , where is the p-wave modulus, should be used. Abeles matrix formalism: The Abeles matrix method is a computationally fast and easy way to calculate the specular reflectivity from a stratified interface, as a function of the perpendicular momentum transfer, Qz. Application

Reflection from a strafied interface Where is the angle of incidence/reflection of the incident radiation and is the wavelength of the radiation. The measured reflectivity depends on the variation in the scattering length density (SLD) profile, ((z)) perpendicular to the interface. Although the scattering length density profile is normally a continuously varying function, the interfacial structure can often be well approximated by a slab model in which layers of thickness (dn), scattering length density (n) and roughness (n,n+1) are sandwiched between the super- and sub-phases. One then uses a refinement procedure to minimise the differences between the theoretical and measured reflectivity curves, by changing the parameters that describe each layer. In this description the interface is split into n layers. Since the incident neutron beam is refracted by each of the layers the wavevector, k, in layer n, is given by:

The Fresnel reflection coefficient between layer n and n+1 is then given by:

Since the interface between each layer is unlikely to be perfectly smooth the roughness/diffuseness of each interface modifies the Fresnel coefficient and is accounted for by an error function, as described by Nevot and Croce (1980).

A phase factor, is introduced, which accounts for the thickness of each layer.

where

. A characteristic matrix, cn is then calculated for each layer.

The resultant matrix is defined as the product of these characteristic matrices

from which the reflectivity is calculated

----------------------------------------------------------------------------------ABCD Matrix: a 2-by-2 matrix describing the effect of an optical element on a laser beam An ABCD matrix or ray matrix [1] is a 2-by-2 matrix associated with an optical element which can be used for describing the element's effect on a laser beam. It can be used both in ray optics, where geometrical rays are propagated, and for propagating Gaussian beams. The paraxial approximation is always required for ABCD matrix calculations, i.e., the involved beam angles or divergence angles must stay small for the calculations to be accurate. Ray Optics Figure 1: Definition of r and before an after an optical system. Originally, the concept was developed for calculating the propagation of geometric rays with some transverse offset r and offset angle from a reference axis (Figure 1). As long as the angles involved are small enough ( paraxial approximation), there is a linear relation between the r and coordinates before and after an optical element. The following equation can then be used for calculating how these parameters are modified by an optical element: where the primed quantities (left-hand side) refer to the beam after passing the optical component. The ABCD matrix is a characteristic of each optical element. For example, a thin lens with focal length f has the following ABCD matrix: This shows that the offset r remains unchanged, whereas the offset angle experiences a change in proportion to r. Propagation through free space over a distance d is associated with the matrix which shows that the angle remains unchanged, whereas the beam offset is increasing or decreasing according to the angle. Further examples for ABCD matrices are given below. For situations where beams propagate through dielectric media, it is convenient to use a modified kind of beam vectors, where the lower component (the angle) is multiplied by the refractive index. This can somewhat simplify the ABCD matrices for certain situations. Propagation of Gaussian Beams: ABCD matrices can also be used for calculating the effect of optical elements on the parameters of a Gaussian beam. A convenient quantity for that purpose is the complex q parameter, which contains information on both the beam radius w and the radius of curvature R of the wavefronts:

The following equation shows how the q parameter is modified by an optical element:

ABCD Matrices of Important Optical Elements: The following list gives the ABCD matrices of frequently used optical elements. Air space with length d: (For propagation in a transparent medium, the length n has to be divided by the refractive index, if the above mentioned modified definition is used where the lower component (the angle) is multiplied by the refractive index.) Lens with focal length f (where positive f applies for a focusing lens):

Curved mirror with curvature radius R (>0 for concave mirror), incidence angle in the horizontal plane:

with Re=Rcos in the tangential plane (horizontal direction) and Re=R/cos in the sagittal plane (vertical direction). Duct:

where the radially varying refractive index is: Combining Multiple Optical Elements: If a beam propagates through several optical elements (including any air spaces in between), this means that the (r ) vector is subsequently multiplied by various matrices. Instead, a single matrix may be used, which is the matrix product of all the single matrices. Note that the first optical element must be on the right-hand side of that product. Typical Applications: Some typical applications of the ABCD matrix algorithm are:

It is often of interest how a laser beam propagates through some optical setup. Both the geometric path of a ray and the evolution of the beam radius can be calculated. The changes of beam parameters within one complete round trip in a resonator can be described with an ABCD matrix. The transverse resonator modes can then be obtained from the matrix components. An extended algorithm, involving an ABCDEF matrix (a 3-by-3 matrix with some constant components), can be used for calculating the alignment sensitivity of a laser resonator [3].

The ABCD matrix method should not be confused with a different matrix method for calculating the reflection and transmission properties of dielectric multilayer coatings.

Praraxial Approximation:

a frequently used approximation, essentially assuming small angular deviations of the propagation directions from some beam axis. Many calculations in optics can be greatly simplified by making the paraxial approximation, i.e. by assuming that the propagation direction of light (e.g. in some laser beam) deviates only slightly from some beam axis. Paraxial Approximation in Geometric Optics: Geometric optics (ray optics) describes light propagation in the form of geometric rays. Here, the paraxial approximation means that the angle between such rays and some reference axis of the optical system always remains small, i.e. <<1 rad. Within that approximation, it can be assumed that tansin. The evolution of beam offset (distance from the reference axis) and beam angle in some optical system can then be described with simple ABCD matrices, because there are linear relations between offset and angle of beams before and after some optical component or system. Paraxial Approximation in Wave Optics: When describing light as a wave phenomenon, the local propagation direction of the energy can be identified with a direction normal to the wavefronts (except in situations with spatial walk-off). If the paraxial approximation holds, i.e. these propagation directions are all close to some reference axis, a second-order differential equation (as obtained from Maxwell's equations) can be replaced with a simple first-order equation. Based on this equation, the formalism of Gaussian beams can be derived, which gives a much simplified understanding of beam propagation and of fundamental limitations such as the minimum beam parameter product. Essentially, the paraxial approximation remains valid as long as divergence angles remain well below 1 rad. This also implies that the beam radius at a beam waist must be much larger than the wavelength. The propagation modes of waveguides, particularly of optical fibers, are also often investigated based on the paraxial approximation. The validity of the analysis is then restricted to cases with a sufficiently large effective mode area and sufficiently small divergence of any beams exiting such a waveguide. The paraxial approximation is very well fulfilled in a wide range of phenomena of laser physics and fiber optics, but it is clearly violated in cases with very strong focusing, where commonly used equations such as =/(w0) for the divergence angle break down. In that regime, polarization issues also demand special care. In particular, polarization components in the propagation direction can occur. For such reasons, the simulation of beam propagation then requires significantly more sophisticated methods. For example, beam propagation methods (propagating a two-dimensional array of complex field amplitudes) can be used which do not need that approximation.

Das könnte Ihnen auch gefallen