You are on page 1of 26

Fresnel diffraction

In optics, the Fresnel diffraction equation


for near-field diffraction is an
approximation of the Kirchhoff–Fresnel
diffraction that can be applied to the
propagation of waves in the near field.[1] It
is used to calculate the diffraction pattern
created by waves passing through an
aperture or around an object, when viewed
from relatively close to the object. In
contrast the diffraction pattern in the far
field region is given by the Fraunhofer
diffraction equation.

The near field can be specified by the


Fresnel number, F of the optical
arrangement. When the diffracted
wave is considered to be in the near field.
However, the validity of the Fresnel
diffraction integral is deduced by the
approximations derived below. Specifically,
the phase terms of third order and higher
must be negligible, a condition that may be
written as
where is the maximal angle described by
, a and L the same as in the
definition of the Fresnel number.

Fresnel diffraction showing central Arago spot

The multiple Fresnel diffraction at closely


spaced periodical ridges (ridged mirror)
causes the specular reflection; this effect
can be used for atomic mirrors.[2]
Early treatments of this
phenomenon
Some of the earliest work on what would
become known as Fresnel diffraction was
carried out by Francesco Maria Grimaldi in
Italy in the 17th century. In his monograph
entitled "Light",[3] Richard C. MacLaurin
explains Fresnel diffraction by asking what
happens when light propagates, and how
that process is affected when a barrier
with a slit or hole in it is interposed in the
beam produced by a distant source of
light. He uses the Principle of Huygens to
investigate, in classical terms, what
transpires. The wave front that proceeds
from the slit and on to a detection screen
some distance away very closely
approximates a wave front originating
across the area of the gap without regard
to any minute interactions with the actual
physical edge.

The result is that if the gap is very narrow


only diffraction patterns with bright
centers can occur. If the gap is made
progressively wider, then diffraction
patterns with dark centers will alternate
with diffraction patterns with bright
centers. As the gap becomes larger, the
differentials between dark and light bands
decrease until a diffraction effect can no
longer be detected.

MacLaurin does not mention the


possibility that the center of the series of
diffraction rings produced when light is
shone through a small hole may be black,
but he does point to the inverse situation
wherein the shadow produced by a small
circular object can paradoxically have a
bright center. (p. 219)

In his Optics,[4] Francis Weston Sears


offers a mathematical approximation
suggested by Fresnel that predicts the
main features of diffraction patterns and
uses only simple mathematics. By
considering the perpendicular distance
from the hole in a barrier screen to a
nearby detection screen along with the
wavelength of the incident light, it is
possible to compute a number of regions
called half-period elements or Fresnel
zones. The inner zone is a circle and each
succeeding zone will be a concentric
annular ring. If the diameter of the circular
hole in the screen is sufficient to expose
the first or central Fresnel zone, the
amplitude of light at the center of the
detection screen will be double what it
would be if the detection screen were not
obstructed. If the diameter of the circular
hole in the screen is sufficient to expose
two Fresnel zones, then the amplitude at
the center is almost zero. That means that
a Fresnel diffraction pattern can have a
dark center. These patterns can be seen
and measured, and correspond well to the
values calculated for them.

The Fresnel diffraction


integral

Diffraction geometry, showing aperture (or diffracting


g y, g p ( g
object) plane and image plane, with coordinate system.

The electric field diffraction pattern at a


point (x, y, z) is given by:

where

is the aperture,

is the wavenumber
is the imaginary unit.

Analytical solution of this integral is


impossible for all but the simplest
diffraction geometries. Therefore, it is
usually calculated numerically.

The Fresnel approximation

The main problem for solving the integral


is the expression of r. First, we can
simplify the algebra by introducing the
substitution:

Substituting into the expression for r, we


find:
Next, using the Taylor series expansion

we can express r as

If we consider all the terms of Taylor


series, then there is no approximation.[5]
Let us substitute this expression in the
argument of the exponential within the
integral; the key to the Fresnel
approximation is to assume that the third
term is very small and can be ignored. In
order to make this possible, it has to
contribute to the variation of the
exponential for an almost null term. In
other words, it has to be much smaller
than the period of the complex
exponential; i.e., :

expressing k in terms of the wavelength,


we get the following relationship:

Multiplying both sides by , we have

or, substituting the earlier expression for


ρ2,
If this condition holds true for all values of
x, x' , y and y' , then we can ignore the third
term in the Taylor expression.
Furthermore, if the third term is negligible,
then all terms of higher order will be even
smaller, so we can ignore them as well.

For applications involving optical


wavelengths, the wavelength λ is typically
many orders of magnitude smaller than
the relevant physical dimensions. In
particular:

and
Thus, as a practical matter, the required
inequality will always hold true as long as

We can then approximate the expression


with only the first two terms:

This equation, then, is the Fresnel


approximation, and the inequality stated
above is a condition for the
approximation's validity.

Fresnel diffraction
The condition for validity is fairly weak,
and it allows all length parameters to take
comparable values, provided the aperture
is small compared to the path length. For
the r in the denominator we go one step
further, and approximate it with only the
first term, . This is valid in particular
if we are interested in the behaviour of the
field only in a small area close to the
origin, where the values of x and y are
much smaller than z. In general, Fresnel
diffraction is valid if the Fresnel number is
approximately 1.

For Fresnel diffraction the electric field at


point (x, y, z) is then given by :
Fresnel diffraction of circular aperture, plotted with
Lommel functions

This is the Fresnel diffraction integral; it


means that, if the Fresnel approximation is
valid, the propagating field is a spherical
wave, originating at the aperture and
moving along z. The integral modulates
the amplitude and phase of the spherical
wave. Analytical solution of this
expression is still only possible in rare
cases. For a further simplified case, valid
only for much larger distances from the
diffraction source, see Fraunhofer
diffraction. Unlike Fraunhofer diffraction,
Fresnel diffraction accounts for the
curvature of the wavefront, in order to
correctly calculate the relative phase of
interfering waves.

Alternative forms
Convolution
The integral can be expressed in other
ways in order to calculate it using some
mathematical properties. If we define the
following function:

then the integral can be expressed in


terms of a convolution:

in other words we are representing the


propagation using a linear-filter modeling.
That is why we might call the function h(x,
y, z) the impulse response of free space
propagation.
Fourier transform

Another possible way is through the


Fourier transform. If in the integral we
express k in terms of the wavelength:

and expand each component of the


transverse displacement:

then we can express the integral in terms


of the two-dimensional Fourier transform.
Let us use the following definition:
where p and q are spatial frequencies
(wavenumbers). The Fresnel integral can
be expressed as

where
That is, first multiply the field to be
propagated by a complex exponential,
calculate its two-dimensional Fourier

transform, replace (p, q) with

and multiply it by another factor. This


expression is better than the others when
the process leads to a known Fourier
transform, and the connection with the
Fourier transform is tightened in the linear
canonical transformation, discussed
below.

Linear canonical
transformation
From the point of view of the linear
canonical transformation, Fresnel
diffraction can be seen as a shear in the
time-frequency domain, corresponding to
how the Fourier transform is a rotation in
the time-frequency domain.

See also
Fraunhofer diffraction
Fresnel integral
Fresnel zone
Fresnel number
Augustin-Jean Fresnel
Ridged mirror
Fresnel imager
Euler spiral

Notes
1. M. Born & E. Wolf, Principles of Optics,
1999, Cambridge University Press,
Cambridge
2.
http://www.ils.uec.ac.jp/~dima/PhysRevLet
t_94_013203.pdf H. Oberst, D. Kouznetsov,
K. Shimizu, J. Fujita, F. Shimizu. Fresnel
diffraction mirror for atomic wave, Physical
Review Letters, 94, 013203 (2005).
3.
https://archive.org/details/lightrichard00ma
clrich Light, by Richard C. MacLaurin, 1909,
Columbia University Press
4. Optics, Francis Weston Sears, p. 248ff,
Addison-Wesley, 1948
5. There was actually an approximation in a
prior step, when assuming is a real
wave. In fact this is not a real solution to the
vector Helmholtz equation, but to the scalar
one. See scalar wave approximation

References
Goodman, Joseph W. (1996).
Introduction to Fourier optics. New York:
McGraw-Hill. ISBN 0-07-024254-2.
Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Fresnel_diffraction&oldid=846833564"

Last edited 5 months ago by Tomro…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like