Hamiltonian optics

Last updated

Hamiltonian optics [1] and Lagrangian optics [2] are two formulations of geometrical optics which share much of the mathematical formalism with Hamiltonian mechanics and Lagrangian mechanics.

Contents

Hamilton's principle

In physics, Hamilton's principle states that the evolution of a system described by generalized coordinates between two specified states at two specified parameters σA and σB is a stationary point (a point where the variation is zero) of the action functional, or

where and is the Lagrangian. Condition is valid if and only if the Euler-Lagrange equations are satisfied, i.e.,

with .

The momentum is defined as

and the Euler–Lagrange equations can then be rewritten as

where .

A different approach to solving this problem consists in defining a Hamiltonian (taking a Legendre transform of the Lagrangian) as

for which a new set of differential equations can be derived by looking at how the total differential of the Lagrangian depends on parameter σ, positions and their derivatives relative to σ. This derivation is the same as in Hamiltonian mechanics, only with time t now replaced by a general parameter σ. Those differential equations are the Hamilton's equations

with . Hamilton's equations are first-order differential equations, while Euler-Lagrange's equations are second-order.

Lagrangian optics

The general results presented above for Hamilton's principle can be applied to optics. [3] [4] In 3D euclidean space the generalized coordinates are now the coordinates of euclidean space.

Fermat's principle

Fermat's principle states that the optical length of the path followed by light between two fixed points, A and B, is a stationary point. It may be a maximum, a minimum, constant or an inflection point. In general, as light travels, it moves in a medium of variable refractive index which is a scalar field of position in space, that is, in 3D euclidean space. Assuming now that light travels along the x3 axis, the path of a light ray may be parametrized as starting at a point and ending at a point . In this case, when compared to Hamilton's principle above, coordinates and take the role of the generalized coordinates while takes the role of parameter , that is, parameter σ =x3 and N=2.

In the context of calculus of variations this can be written as [2]

where ds is an infinitesimal displacement along the ray given by and

is the optical Lagrangian and .

The optical path length (OPL) is defined as

where n is the local refractive index as a function of position along the path between points A and B.

The Euler-Lagrange equations

The general results presented above for Hamilton's principle can be applied to optics using the Lagrangian defined in Fermat's principle. The Euler-Lagrange equations with parameter σ =x3 and N=2 applied to Fermat's principle result in

with k = 1, 2 and where L is the optical Lagrangian and .

Optical momentum

The optical momentum is defined as

and from the definition of the optical Lagrangian this expression can be rewritten as

Optical momentum Hamiltonian Optics-Optical Momentum.png
Optical momentum

or in vector form

where is a unit vector and angles α1, α2 and α3 are the angles p makes to axis x1, x2 and x3 respectively, as shown in figure "optical momentum". Therefore, the optical momentum is a vector of norm

where n is the refractive index at which p is calculated. Vector p points in the direction of propagation of light. If light is propagating in a gradient index optic the path of the light ray is curved and vector p is tangent to the light ray.

The expression for the optical path length can also be written as a function of the optical momentum. Having in consideration that the expression for the optical Lagrangian can be rewritten as

and the expression for the optical path length is

Hamilton's equations

Similarly to what happens in Hamiltonian mechanics, also in optics the Hamiltonian is defined by the expression given above for N = 2 corresponding to functions and to be determined

Comparing this expression with for the Lagrangian results in

And the corresponding Hamilton's equations with parameter σ =x3 and k=1,2 applied to optics are [5] [6]

with and .

Applications

It is assumed that light travels along the x3 axis, in Hamilton's principle above, coordinates and take the role of the generalized coordinates while takes the role of parameter , that is, parameter σ =x3 and N=2.

Refraction and reflection

If plane x1x2 separates two media of refractive index nA below and nB above it, the refractive index is given by a step function

and from Hamilton's equations

and therefore or for k = 1, 2.

An incoming light ray has momentum pA before refraction (below plane x1x2) and momentum pB after refraction (above plane x1x2). The light ray makes an angle θA with axis x3 (the normal to the refractive surface) before refraction and an angle θB with axis x3 after refraction. Since the p1 and p2 components of the momentum are constant, only p3 changes from p3A to p3B.

Refraction Hamiltonian Optics-Refraction.png
Refraction

Figure "refraction" shows the geometry of this refraction from which . Since and , this last expression can be written as

which is Snell's law of refraction.

In figure "refraction", the normal to the refractive surface points in the direction of axis x3, and also of vector . A unit normal to the refractive surface can then be obtained from the momenta of the incoming and outgoing rays by

where i and r are unit vectors in the directions of the incident and refracted rays. Also, the outgoing ray (in the direction of ) is contained in the plane defined by the incoming ray (in the direction of ) and the normal to the surface.

A similar argument can be used for reflection in deriving the law of specular reflection, only now with nA=nB, resulting in θA=θB. Also, if i and r are unit vectors in the directions of the incident and refracted ray respectively, the corresponding normal to the surface is given by the same expression as for refraction, only with nA=nB

In vector form, if i is a unit vector pointing in the direction of the incident ray and n is the unit normal to the surface, the direction r of the refracted ray is given by: [3]

with

If in<0 then −n should be used in the calculations. When , light suffers total internal reflection and the expression for the reflected ray is that of reflection:

Rays and wavefronts

From the definition of optical path length

Rays and wavefronts Hamiltonian Optics-Rays and Wavefronts.svg
Rays and wavefronts

with k=1,2 where the Euler-Lagrange equations with k=1,2 were used. Also, from the last of Hamilton's equations and from above

combining the equations for the components of momentum p results in

Since p is a vector tangent to the light rays, surfaces S=Constant must be perpendicular to those light rays. These surfaces are called wavefronts. Figure "rays and wavefronts" illustrates this relationship. Also shown is optical momentum p, tangent to a light ray and perpendicular to the wavefront.

Vector field is conservative vector field. The gradient theorem can then be applied to the optical path length (as given above) resulting in

and the optical path length S calculated along a curve C between points A and B is a function of only its end points A and B and not the shape of the curve between them. In particular, if the curve is closed, it starts and ends at the same point, or A=B so that

This result may be applied to a closed path ABCDA as in figure "optical path length"

Optical path length Hamiltonian Optics-Optical Path Length.png
Optical path length

for curve segment AB the optical momentum p is perpendicular to a displacement ds along curve AB, or . The same is true for segment CD. For segment BC the optical momentum p has the same direction as displacement ds and . For segment DA the optical momentum p has the opposite direction to displacement ds and . However inverting the direction of the integration so that the integral is taken from A to D, ds inverts direction and . From these considerations

or

and the optical path length SBC between points B and C along the ray connecting them is the same as the optical path length SAD between points A and D along the ray connecting them. The optical path length is constant between wavefronts.

Phase space

Figure "2D phase space" shows at the top some light rays in a two-dimensional space. Here x2=0 and p2=0 so light travels on the plane x1x3 in directions of increasing x3 values. In this case and the direction of a light ray is completely specified by the p1 component of momentum since p2=0. If p1 is given, p3 may be calculated (given the value of the refractive index n) and therefore p1 suffices to determine the direction of the light ray. The refractive index of the medium the ray is traveling in is determined by .

2D phase space Hamiltonian Optics-2D Phase Space.png
2D phase space

For example, ray rC crosses axis x1 at coordinate xB with an optical momentum pC, which has its tip on a circle of radius n centered at position xB. Coordinate xB and the horizontal coordinate p1C of momentum pC completely define ray rC as it crosses axis x1. This ray may then be defined by a point rC=(xB,p1C) in space x1p1 as shown at the bottom of the figure. Space x1p1 is called phase space and different light rays may be represented by different points in this space.

As such, ray rD shown at the top is represented by a point rD in phase space at the bottom. All rays crossing axis x1 at coordinate xB contained between rays rC and rD are represented by a vertical line connecting points rC and rD in phase space. Accordingly, all rays crossing axis x1 at coordinate xA contained between rays rA and rB are represented by a vertical line connecting points rA and rB in phase space. In general, all rays crossing axis x1 between xL and xR are represented by a volume R in phase space. The rays at the boundary ∂R of volume R are called edge rays. For example, at position xA of axis x1, rays rA and rB are the edge rays since all other rays are contained between these two. (A ray parallel to x1 would not be between the two rays, since the momentum is not in-between the two rays)

In three-dimensional geometry the optical momentum is given by with . If p1 and p2 are given, p3 may be calculated (given the value of the refractive index n) and therefore p1 and p2 suffice to determine the direction of the light ray. A ray traveling along axis x3 is then defined by a point (x1,x2) in plane x1x2 and a direction (p1,p2). It may then be defined by a point in four-dimensional phase space x1x2p1p2.

Conservation of etendue

Figure "volume variation" shows a volume V bound by an area A. Over time, if the boundary A moves, the volume of V may vary. In particular, an infinitesimal area dA with outward pointing unit normal n moves with a velocity v.

Volume variation Hamiltonian Optics-Volume Variation.png
Volume variation

This leads to a volume variation . Making use of Gauss's theorem, the variation in time of the total volume V volume moving in space is

The rightmost term is a volume integral over the volume V and the middle term is the surface integral over the boundary A of the volume V. Also, v is the velocity with which the points in V are moving.

In optics coordinate takes the role of time. In phase space a light ray is identified by a point which moves with a "velocity" where the dot represents a derivative relative to . A set of light rays spreading over in coordinate , in coordinate , in coordinate and in coordinate occupies a volume in phase space. In general, a large set of rays occupies a large volume in phase space to which Gauss's theorem may be applied

and using Hamilton's equations

or and which means that the phase space volume is conserved as light travels along an optical system.

The volume occupied by a set of rays in phase space is called etendue, which is conserved as light rays progress in the optical system along direction x3. This corresponds to Liouville's theorem, which also applies to Hamiltonian mechanics.

However, the meaning of Liouville’s theorem in mechanics is rather different from the theorem of conservation of étendue. Liouville’s theorem is essentially statistical in nature, and it refers to the evolution in time of an ensemble of mechanical systems of identical properties but with different initial conditions. Each system is represented by a single point in phase space, and the theorem states that the average density of points in phase space is constant in time. An example would be the molecules of a perfect classical gas in equilibrium in a container. Each point in phase space, which in this example has 2N dimensions, where N is the number of molecules, represents one of an ensemble of identical containers, an ensemble large enough to permit taking a statistical average of the density of representative points. Liouville’s theorem states that if all the containers remain in equilibrium, the average density of points remains constant. [3]

Imaging and nonimaging optics

Figure "conservation of etendue" shows on the left a diagrammatic two-dimensional optical system in which x2=0 and p2=0 so light travels on the plane x1x3 in directions of increasing x3 values.

Conservation of etendue Hamiltonian Optics-Conservation of Etendue.png
Conservation of etendue

Light rays crossing the input aperture of the optic at point x1=xI are contained between edge rays rA and rB represented by a vertical line between points rA and rB at the phase space of the input aperture (right, bottom corner of the figure). All rays crossing the input aperture are represented in phase space by a region RI.

Also, light rays crossing the output aperture of the optic at point x1=xO are contained between edge rays rA and rB represented by a vertical line between points rA and rB at the phase space of the output aperture (right, top corner of the figure). All rays crossing the output aperture are represented in phase space by a region RO.

Conservation of etendue in the optical system means that the volume (or area in this two-dimensional case) in phase space occupied by RI at the input aperture must be the same as the volume in phase space occupied by RO at the output aperture.

In imaging optics, all light rays crossing the input aperture at x1=xI are redirected by it towards the output aperture at x1=xO where xI=m xO. This ensures that an image of the input is formed at the output with a magnification m. In phase space, this means that vertical lines in the phase space at the input are transformed into vertical lines at the output. That would be the case of vertical line rArB in RI transformed to vertical line rArB in RO.

In nonimaging optics, the goal is not to form an image but simply to transfer all light from the input aperture to the output aperture. This is accomplished by transforming the edge rays ∂RI of RI to edge rays ∂RO of RO. This is known as the edge ray principle.

Generalizations

Above it was assumed that light travels along the x3 axis, in Hamilton's principle above, coordinates and take the role of the generalized coordinates while takes the role of parameter , that is, parameter σ =x3 and N=2. However, different parametrizations of the light rays are possible, as well as the use of generalized coordinates.

General ray parametrization

A more general situation can be considered in which the path of a light ray is parametrized as in which σ is a general parameter. In this case, when compared to Hamilton's principle above, coordinates , and take the role of the generalized coordinates with N=3. Applying Hamilton's principle to optics in this case leads to

where now and and for which the Euler-Lagrange equations applied to this form of Fermat's principle result in

with k=1,2,3 and where L is the optical Lagrangian. Also in this case the optical momentum is defined as

and the Hamiltonian P is defined by the expression given above for N=3 corresponding to functions , and to be determined

And the corresponding Hamilton's equations with k=1,2,3 applied optics are

with and .

The optical Lagrangian is given by

and does not explicitly depend on parameter σ. For that reason not all solutions of the Euler-Lagrange equations will be possible light rays, since their derivation assumed an explicit dependence of L on σ which does not happen in optics.

The optical momentum components can be obtained from

where . The expression for the Lagrangian can be rewritten as

Comparing this expression for L with that for the Hamiltonian P it can be concluded that

From the expressions for the components of the optical momentum results

The optical Hamiltonian is chosen as

although other choices could be made. [3] [4] The Hamilton's equations with k = 1, 2, 3 defined above together with define the possible light rays.

Generalized coordinates

As in Hamiltonian mechanics, it is also possible to write the equations of Hamiltonian optics in terms of generalized coordinates , generalized momenta and Hamiltonian P as [3] [4]

where the optical momentum is given by

and , and are unit vectors. A particular case is obtained when these vectors form an orthonormal basis, that is, they are all perpendicular to each other. In that case, is the cosine of the angle the optical momentum makes to unit vector .

See also

Related Research Articles

<span class="mw-page-title-main">Lorentz force</span> Force acting on charged particles in electric and magnetic fields

In physics, the Lorentz force is the combination of electric and magnetic force on a point charge due to electromagnetic fields. A particle of charge q moving with a velocity v in an electric field E and a magnetic field B experiences a force of

In vector calculus and differential geometry the generalized Stokes theorem, also called the Stokes–Cartan theorem, is a statement about the integration of differential forms on manifolds, which both simplifies and generalizes several theorems from vector calculus. In particular, the fundamental theorem of calculus is the special case where the manifold is a line segment, Green’s theorem and Stokes' theorem are the cases of a surface in or and the divergence theorem is the case of a volume in Hence, the theorem is sometimes referred to as the Fundamental Theorem of Multivariate Calculus.

<span class="mw-page-title-main">Fokker–Planck equation</span> Partial differential equation

In statistical mechanics and information theory, the Fokker–Planck equation is a partial differential equation that describes the time evolution of the probability density function of the velocity of a particle under the influence of drag forces and random forces, as in Brownian motion. The equation can be generalized to other observables as well. The Fokker-Planck equation has multiple applications in information theory, graph theory, data science, finance, economics etc.

The calculus of variations is a field of mathematical analysis that uses variations, which are small changes in functions and functionals, to find maxima and minima of functionals: mappings from a set of functions to the real numbers. Functionals are often expressed as definite integrals involving functions and their derivatives. Functions that maximize or minimize functionals may be found using the Euler–Lagrange equation of the calculus of variations.

In continuum mechanics, the infinitesimal strain theory is a mathematical approach to the description of the deformation of a solid body in which the displacements of the material particles are assumed to be much smaller than any relevant dimension of the body; so that its geometry and the constitutive properties of the material at each point of space can be assumed to be unchanged by the deformation.

A continuity equation or transport equation is an equation that describes the transport of some quantity. It is particularly simple and powerful when applied to a conserved quantity, but it can be generalized to apply to any extensive quantity. Since mass, energy, momentum, electric charge and other natural quantities are conserved under their respective appropriate conditions, a variety of physical phenomena may be described using continuity equations.

A directional derivative is a concept in multivariable calculus that measures the rate at which a function changes in a particular direction at a given point.

In mathematics, the method of characteristics is a technique for solving partial differential equations. Typically, it applies to first-order equations, although more generally the method of characteristics is valid for any hyperbolic partial differential equation. The method is to reduce a partial differential equation to a family of ordinary differential equations along which the solution can be integrated from some initial data given on a suitable hypersurface.

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

In mechanics, virtual work arises in the application of the principle of least action to the study of forces and movement of a mechanical system. The work of a force acting on a particle as it moves along a displacement is different for different displacements. Among all the possible displacements that a particle may follow, called virtual displacements, one will minimize the action. This displacement is therefore the displacement followed by the particle according to the principle of least action.

The work of a force on a particle along a virtual displacement is known as the virtual work.

<span class="mw-page-title-main">Charge density</span> Electric charge per unit length, area or volume

In electromagnetism, charge density is the amount of electric charge per unit length, surface area, or volume. Volume charge density is the quantity of charge per unit volume, measured in the SI system in coulombs per cubic meter (C⋅m−3), at any point in a volume. Surface charge density (σ) is the quantity of charge per unit area, measured in coulombs per square meter (C⋅m−2), at any point on a surface charge distribution on a two dimensional surface. Linear charge density (λ) is the quantity of charge per unit length, measured in coulombs per meter (C⋅m−1), at any point on a line charge distribution. Charge density can be either positive or negative, since electric charge can be either positive or negative.

The Cauchy momentum equation is a vector partial differential equation put forth by Cauchy that describes the non-relativistic momentum transport in any continuum.

<span class="mw-page-title-main">Interval finite element</span>

In numerical analysis, the interval finite element method is a finite element method that uses interval parameters. Interval FEM can be applied in situations where it is not possible to get reliable probabilistic characteristics of the structure. This is important in concrete structures, wood structures, geomechanics, composite structures, biomechanics and in many other areas. The goal of the Interval Finite Element is to find upper and lower bounds of different characteristics of the model and use these results in the design process. This is so called worst case design, which is closely related to the limit state design.

The Clausius–Duhem inequality is a way of expressing the second law of thermodynamics that is used in continuum mechanics. This inequality is particularly useful in determining whether the constitutive relation of a material is thermodynamically allowable.

<span class="mw-page-title-main">Stokes' theorem</span> Theorem in vector calculus

Stokes' theorem, also known as the Kelvin–Stokes theorem after Lord Kelvin and George Stokes, the fundamental theorem for curls or simply the curl theorem, is a theorem in vector calculus on . Given a vector field, the theorem relates the integral of the curl of the vector field over some surface, to the line integral of the vector field around the boundary of the surface. The classical theorem of Stokes can be stated in one sentence: The line integral of a vector field over a loop is equal to the flux of its curl through the enclosed surface. It is illustrated in the figure, where the direction of positive circulation of the bounding contour ∂Σ, and the direction n of positive flux through the surface Σ, are related by a right-hand-rule. For the right hand the fingers circulate along ∂Σ and the thumb is directed along n.

<span class="mw-page-title-main">Relativistic Lagrangian mechanics</span> Mathematical formulation of special and general relativity

In theoretical physics, relativistic Lagrangian mechanics is Lagrangian mechanics applied in the context of special relativity and general relativity.

<span class="mw-page-title-main">Objective stress rate</span>

In continuum mechanics, objective stress rates are time derivatives of stress that do not depend on the frame of reference. Many constitutive equations are designed in the form of a relation between a stress-rate and a strain-rate. The mechanical response of a material should not depend on the frame of reference. In other words, material constitutive equations should be frame-indifferent (objective). If the stress and strain measures are material quantities then objectivity is automatically satisfied. However, if the quantities are spatial, then the objectivity of the stress-rate is not guaranteed even if the strain-rate is objective.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

The streamline upwind Petrov–Galerkin pressure-stabilizing Petrov–Galerkin formulation for incompressible Navier–Stokes equations can be used for finite element computations of high Reynolds number incompressible flow using equal order of finite element space by introducing additional stabilization terms in the Navier–Stokes Galerkin formulation.

In physics and mathematics, the Klein–Kramers equation or sometimes referred as Kramers–Chandrasekhar equation is a partial differential equation that describes the probability density function f of a Brownian particle in phase space (r, p). It is a special case of the Fokker–Planck equation.

References

  1. H. A. Buchdahl, An Introduction to Hamiltonian Optics, Dover Publications, 1993, ISBN   978-0486675978.
  2. 1 2 Vasudevan Lakshminarayanan et al., Lagrangian Optics, Springer Netherlands, 2011, ISBN   978-0792375821.
  3. 1 2 3 4 5 Chaves, Julio (2015). Introduction to Nonimaging Optics, Second Edition. CRC Press. ISBN   978-1482206739.
  4. 1 2 3 Roland Winston et al., Nonimaging Optics, Academic Press, 2004, ISBN   978-0127597515.
  5. Dietrich Marcuse, Light Transmission Optics, Van Nostrand Reinhold Company, New York, 1972, ISBN   978-0894643057.
  6. Rudolf Karl Luneburg,Mathematical Theory of Optics, University of California Press, Berkeley, CA, 1964, p. 90.