Jump back to chapter selection.


Table of Contents

5.1 Plane Waves and the Helmholtz Equation
5.2 Paraxial Approximation
5.3 Fresnel Approximation
5.4 The Fraunhofer Limit: Far Field
5.5 Diffraction Patterns -  Amplitude Modulation
5.6 Fourier Optics with a Lens
5.7 Holography
5.8 Paraxial Ray Optics


5 Fourier Optics

In this chapter, we will use a plane wave expansion of a monochromatic field to study light propagation through an optical system. The simplest of these systems is free space. It will soon become clear why this chapter is specifically titled 'Fourier' Optics.


5.1 Plane Waves and the Helmholtz Equation

As we have seen previously, an arbitrary function may often be constructed from a sum or integral of harmonic functions (plane waves) of different frequencies and complex amplitudes. This principle extends to multiple dimensions: an arbitrary spatial function f(x,y), representing for instance a field distribution in a plane, may be constructed as a superposition of harmonic functions with different spatial frequencies (kx,ky) and complex amplitudes:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 6.webp|700

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/QE_script 14.webp|700

In the figures above, we describe the optical wave with a scalar function U(x,y,z), which could represent, for example, one Cartesian component of the electric field. The problem at hand is the following: We consider the transmission of an optical wave U(x,y,z) through an optical system, which is assumed to be linear. The input field is defined in an initial plane, say U(x,y,0), and we wish to find the field in an output plane, U(x,y,zd), after propagation through a distance zd:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 7.webp|700

As described in more detail here, a linear system is characterised by its impulse response or, equivalently, by its response to a harmonic function (its transfer function).

We begin the discussion of Fourier optics by recalling the wave equation for the electric field in a source-free, homogeneous, linear, isotropic, and non-dispersive medium (as derived in Chapter 1):

2En2c22t2E=0.

For a monochromatic wave, we may write the electric field as E(r,t)=Re[e^0U(r)eiωt], where e^0 is a constant unit vector indicating polarisation, U(r) is a complex scalar function of space representing the complex amplitude (magnitude and phase) of the field component, and eiωt is the time-harmonic factor (using the physics convention, where often eiωt is used for time evolution in engineering). Substituting this into the wave equation yields the Helmholtz equation for the spatial part U(r):

2U(r)+k2U(r)=0,

with k=nω/c being the wave number in the medium. The time-averaged intensity is then obtained as I(r)=12Z|U(r)|2, where Z is the impedance of the medium. For the previously discussed monochromatic plane waves, |U(r)| is constant (independent of r). However, the Helmholtz equation also describes beams where the intensity is not uniform in space, such as Gaussian beams.
For such waves, we define the notion of a wavefront: a wavefront is a surface of constant phase. That is, if ϕ(r)=arg[U(r)], a wavefront is a 2-D surface on which ϕ(r) is constant (modulo 2π).

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 8.webp|700

We can see that the wavefronts change curvature upon propagation in a Gaussian beam. For a plane wave, the wavefronts are planes. As we may expect, these wavefronts bend when passing through optical components, such as lenses:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 9.webp|700


5.2 Paraxial Approximation

In the paraxial approximation, we assume that the light rays (normals to the wavefronts) form only small angles with the principal axis of propagation (conventionally the z-axis). Therefore, the transverse components of the wavevector, kx and ky, which encode deviations from propagation straight along the z-axis, are assumed to be much smaller than the total wave number k. This approximation is valid for beams with small divergence angles, as is often the case for the output of a laser.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 10.webp|700

For a field U(x,y,z) at a fixed plane z, we can define its two-dimensional spatial Fourier transform V(kx,ky,z) with respect to the transverse coordinates x and y:

V(kx,ky,z)=U(x,y,z)ei(kxx+kyy)dxdy.

The inverse transformation is:

U(x,y,z)=1(2π)2V(kx,ky,z)ei(kxx+kyy)dkxdky.

Each component (kx,ky) in this expansion corresponds to a plane wave whose wavevector is k=kxx^+kyy^+kzz^, where kz is determined by the Helmholtz equation: kx2+ky2+kz2=k2=(nω/c)2. Thus, kz=k2kx2ky2.
The paraxial approximation can be formally stated as kx2+ky2k2.

Applying the 2D Fourier transform (with respect to x,y) to the Helmholtz equation 2U+k2U=0, where 2=2/x2+2/y2+2/z2:
Fxy{2U/x2}=(ikx)2V=kx2V. Similarly for y.
So, we obtain an ordinary differential equation for V(kx,ky,z) with respect to z:

kx2Vky2V+2Vz2+k2V=02Vz2+(k2kx2ky2)V=0.

Recognising kz2=k2kx2ky2, this is 2Vz2+kz2V=0.
The solution for forward propagation in z is:

V(kx,ky,z)=V(kx,ky,0)eikzz=V(kx,ky,0)eizk2kx2ky2.

We define the transfer function of free space propagation over distance z as:

H(kx,ky,z)=eizk2kx2ky2.

Thus, V(kx,ky,z)=V(kx,ky,0)H(kx,ky,z). Note that kz depends on kx and ky, so H is not dependent on an independent kz variable but rather on kx,ky,k, and z. This approach is valid for propagation in any homogeneous isotropic medium.
Knowing U(x,y,0) in an initial plane allows us to find its profile U(x,y,z) in any other plane at a distance z:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 11.webp|700

The procedure is:

  1. Calculate the 2D spatial Fourier transform V(kx,ky,0) of the input field U(x,y,0).
  2. Multiply V(kx,ky,0) by the transfer function H(kx,ky,z) to obtain V(kx,ky,z).
  3. Apply the inverse 2D spatial Fourier transform to V(kx,ky,z) to obtain U(x,y,z).

It is important to remember that we are working with linear systems: harmonic components (spatial frequencies kx,ky) are not created or destroyed during propagation in a linear homogeneous medium; only their relative phases are modified by H(kx,ky,z).


5.3 Fresnel Approximation

In the expression for the transfer function H(kx,ky,z), the term kz=k2kx2ky2 makes analytical inverse Fourier transformation difficult. In the paraxial approximation (kx2+ky2k2), the angles θxkx/k and θyky/k of the constituent plane waves with respect to the z-axis are small. We can then expand the phase of the transfer function, kzz:

kzz=zk2kx2ky2=kz1kx2+ky2k2.

Using the Taylor expansion 1u1u/2u2/8 for small u:

kzzkz(1kx2+ky22k218(kx2+ky2k2)2).

The Fresnel approximation (or paraxial wave equation approximation) consists of keeping terms only up to the first order in u=(kx2+ky2)/k2 within the square root, which means keeping terms quadratic in kx,ky:

kzk(1kx2+ky22k2)=kkx2+ky22k.

This approximates the spherical wavefront segment of each plane wave component with a parabolic one. The simplified transfer function becomes:

H(kx,ky,z)ei(kkx2+ky22k)z=eikzeizkx2+ky22k.

We can write H0=eikz, which is the phase accumulated by a plane wave propagating along z.
The Fresnel approximation implies that we are observing the field at a distance z that is large compared to the transverse extent of the source/aperture, but not so large that the wavefronts become essentially planar over the observation region (which leads to Fraunhofer diffraction).

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/QE_script 15.webp|700

Considering the source of the wave to be spherical, the Fresnel approximation approximates these spherical wavefronts with parabolas:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 12.webp|700

The validity of the Fresnel approximation requires that the next term in the Taylor expansion of the phase, kz18(kx2+ky2k2)2, must be much less than π. This is a stricter condition than just kx2+ky2k2. It can be related to the Fresnel number NF=a2/(λz), where a is a characteristic transverse dimension (of aperture or beam).

Next, we find the impulse response h(x,y,z) for Fresnel propagation, which is the field at (x,y,z) due to a point source U(x,y,0)=δ(x)δ(y).

  1. The 2D Fourier transform of U(x,y,0)=δ(x)δ(y) is V(kx,ky,0)=δ(x)δ(y)ei(kxx+kyy)dxdy=1.
  2. Therefore, V(kx,ky,z)=1H(kx,ky,z)=eikzeizkx2+ky22k.
  3. We take the inverse Fourier transform:h(x,y,z)=U(x,y,z)=eikz(2π)2eizkx2+ky22kei(kxx+kyy)dkxdky.

We find that the impulse response function h(x,y,z) is given by:

h(x,y,z)=eikziλzeik2z(x2+y2)=k2πizeikzeik2z(x2+y2).

This impulse response is useful because for a linear system, the output U(x,y,z) for an arbitrary input U(x,y,0) is given by the convolution:

U(x,y,z)=h(xx,yy,z)U(x,y,0)dxdy.

Substituting h(xx,yy,z):

U(x,y,z)=eikziλzU(x,y,0)eik2z[(xx)2+(yy)2]dxdy.

This is the Fresnel diffraction integral.

The general steps to find the electric field U(x,y,z) for a given input U(x,y,0) are:

  1. Find the 2D spatial Fourier transform V(kx,ky,0) of U(x,y,0).
  2. Multiply by the appropriate transfer function H(kx,ky,z) to obtain V(kx,ky,z).
  3. Apply the inverse 2D spatial Fourier transform to V(kx,ky,z) to obtain U(x,y,z).

5.4 The Fraunhofer Limit: Far Field

The Fraunhofer approximation, or far-field diffraction, is a limit of the Fresnel approximation valid at sufficiently large distances z from an aperture or object of characteristic transverse size Dobj. It requires the Fresnel conditions to be met, plus an even stronger condition.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 14.webp|700

The condition for Fraunhofer diffraction is often expressed using the Fresnel number NF=Dobj2/(λz). The Fraunhofer regime applies when zDobj2/λ, which means NF1.
If Dobs is the size of the observation region (detector), then we also typically require zDobs2/λ.

The crucial simplification in the Fraunhofer limit arises from approximating the quadratic phase term in the Fresnel integral eik2z[(xx)2+(yy)2].
We expand (xx)2=x22xx+x2. The term eik2z(x2+y2) is kept under the integral. The term eik2z(x2+y2) is taken outside. The Fraunhofer condition allows k(x2+y2)max/(2z)π.
The field U(x,y,z) in the Fraunhofer regime becomes:

U(x,y,z)=eikziλzeik2z(x2+y2)U(x,y,0)eikz(xx+yy)dxdy.

Recognising kx=kx/z=2πx/(λz) and ky=ky/z=2πy/(λz) as spatial frequencies (proportional to observation angles θxx/z,θyy/z), the integral is the 2D Fourier transform of the input field U(x,y,0) evaluated at these spatial frequencies:

U(x,y,z)eikzeik2z(x2+y2)V(kx=kxz,ky=kyz;at z=0).

The Fraunhofer approximation essentially states that in the far field, the observed complex amplitude is proportional to the Fourier transform of the aperture distribution, multiplied by a phase factor. All plane wave components effectively interfere constructively only along specific directions corresponding to their (kx,ky) values, mapping spatial frequencies in the object plane to positions in the observation plane.


5.5 Diffraction Patterns - Amplitude Modulation

When an optical wave passes through an aperture or is otherwise spatially modulated in amplitude and/or phase, and then propagates some distance in free space, the resulting intensity distribution is called a diffraction pattern. From the discussion above, it should be clear that simply expecting the intensity pattern to be a geometric shadow of the aperture is an oversimplification, valid only in the limit of ray optics where the wave nature of light is ignored.
Consider an aperture described by an aperture function p(x,y) in the input plane z=0:

p(x,y)={1inside the aperture0outside the aperture.

If an incident wave Uinc(x,y) illuminates this aperture, the field immediately after the aperture is U(x,y,0)=Uinc(x,y)p(x,y). We now know, in principle, how to obtain the field U(x,y,z) at some distance z.

5.5.1 Rectangular Aperture

Consider a rectangular aperture of width Dx and height Dy, centered at the origin. Assume it is illuminated by a normally incident plane wave of constant amplitude Ein (so U(x,y,0)=Ein inside the aperture and 0 outside). We apply the Fraunhofer approximation to find the far-field pattern at a distance d.
The 2D Fourier transform V(kx,ky) of U(x,y,0) is:

V(kx,ky)=EinDx/2Dx/2eikxxdxDy/2Dy/2eikyydy=Ein[eikxxikx]Dx/2Dx/2[eikyyiky]Dy/2Dy/2=Ein(2sin(kxDx/2)kx)(2sin(kyDy/2)ky)=EinDxDysinc(kxDx2π)sinc(kyDy2π),

where sinc(u)=sin(πu)/(πu). The far-field intensity I(xs,ys,d)|V(kx=kxs/d,ky=kys/d)|2. If I0 is related to Ein2:

Iscreen(xs,ys,d)=Ipeak[sinc(kDxxs2d)]2[sinc(kDyys2d)]2=Ipeak[sinc(πDxxsλd)]2[sinc(πDyysλd)]2.

Here Ipeak is the intensity at the centre of the pattern. Then, we arrive at:

Iscreen(x,y,d)=Iaperture(DxDyλd)2sinc2(πDxxλd)sinc2(πDyyλd).

This result is expected, as the Fourier transform of a rectangular function (top-hat) is a sinc function. The intensity pattern is shown next:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 17.webp|700

5.5.2 Circular Aperture

Consider a circular aperture of diameter D. The 2D Fourier transform of a circular aperture (circ function) is related to a Bessel function of the first kind, J1:

V(kρ)=Einπ(D/2)22J1(kρD/2)kρD/2,

where kρ=kx2+ky2. The far-field intensity pattern at radius ρs=xs2+ys2 on the screen is:

Iscreen(ρs,d)=Iaperture(πD24λd)2[2J1(kDρs/(2d))kDρs/(2d)]2=Iaperture(πD24λd)2[2J1(πDρs/(λd))πDρs/(λd)]2.

This is the characteristic Airy disk pattern shown in the next figure.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 18.webp|700


5.6 Fourier Optics with a Lens

A thin lens introduces a quadratic phase transformation to an incident wavefront. For a lens with focal length f, its phase transfer function is tL(x,y)=eik2f(x2+y2) (for a focusing lens, assuming it's thin and located at z=0).
If an object Uin(xo,yo) is placed at the front focal plane (z=f) of a lens, the field at the back focal plane (z=f) UBFP(xf,yf) is proportional to the Fourier transform of Uin:

UBFP(xf,yf)F{Uin(xo,yo)}|kx=kxf/f,ky=kyf/f.

More generally, if an object U(xo,yo) is placed immediately before a lens, the field at its back focal plane is proportional to the Fourier Transform of U(xo,yo):

U(xf,yf,f)=eikfiλfeik2f(xf2+yf2)U(xo,yo,0)eikf(xoxf+yoyf)dxodyo.

The integral is V(kx=kxf/f,ky=kyf/f), where V is the FT of U(xo,yo,0) with the appropriate FT kernel sign. This demonstrates the Fourier transforming property of a lens.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 19.webp|700

A common configuration is a 2f system, where an object is placed at distance f before a lens, and the image (which is the Fourier transform) is observed at distance f after the lens. If another identical lens is placed at 2f, it performs an inverse Fourier transform, potentially forming an inverted image of the original object at 4f from the first object plane. This is a 4f system:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 20.webp|700

In the Fourier plane (at 2f from the object, between the two lenses), one can place an amplitude or phase mask P(xF,yF) to filter out or modify specific spatial frequency components of the object. Here xF,yF are coordinates in the Fourier plane, related to spatial frequencies kx,ky by kxkxF/f,kykyF/f. The transfer function of such a spatial filtering system is effectively H(kx,ky)P(xF=λfkx/(2π),yF=λfky/(2π)).
Consider the image example: A standard circular aperture in the Fourier plane acts as a low-pass filter, blurring the image by removing high spatial frequencies (sharp details). An opaque stop in the centre acts as a high-pass filter, enhancing edges and removing large-scale variations, making the man's skin appear dark while highlighting hair.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 21.webp|700

A Fresnel zone plate is another optical element that can focus light, but it operates based on diffraction rather than refraction.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 22.webp|700

Its transmission function t(x,y) consists of concentric transparent and opaque zones:

t(x,y)={1,for cos(π(x2+y2)λf)>00,otherwise

for a binary zone plate designed for focal length f.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 23.webp|700

The spacing of these Fresnel zones is such that light diffracted from the transparent zones interferes constructively at the desired focal point. Zone plates are inherently chromatic, focusing different wavelengths to different focal points.


5.7 Holography

Holograms are recordings that encode the full optical wave from an object, including both its amplitude and phase information. In principle, if we could create a transparency t(x,y) equal to the complex field U(x,y,0) from an object, illuminating this transparency with a plane wave would reconstruct the object wave. However, optical detectors are sensitive only to intensity (|U|2), not directly to phase. Holography overcomes this by interfering a reference wave Ur with the object wave Uo.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/Pasted image 20250529193917.png|700
(Image source)

If these two waves overlap at z=0, the intensity pattern recorded on a film (hologram) is:

Irec=|Uo+Ur|2=|Ur|2+|Uo|2+UrUo+UrUo.

Assuming the film's amplitude transmittance t after development is proportional to Irec:

tIr+Io+UrUo+UrUo,

where Ir=|Ur|2 and Io=|Uo|2.
When this hologram is illuminated by the original reference wave Ur, the transmitted field Utrans=tUr is:

UtransUrIr+UrIo+|Ur|2Uo+Ur2Uo.

If the reference wave Ur is a uniform plane wave with amplitude Ar (so Ir=Ar2), the terms are:

  1. ArIr: A scaled version of the reference wave (DC term).
  2. ArIo(x,y): The reference wave modulated by the object intensity (distorted image).
  3. IrUo(x,y): A wave proportional to the original object wave Uo. This term reconstructs the virtual image of the object.
  4. Ar2UrUo: A wave related to the conjugate of the object wave. This term reconstructs the real image (often called the twin image or conjugate image).

To separate these four waves spatially, a common technique is off-axis holography, where the reference wave Ur and object wave Uo are incident on the recording film at a significant angle to each other:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/QE_script 18.webp|700

This angular separation ensures that upon reconstruction, the four terms propagate in different directions, allowing the desired Uo term to be viewed without overlap from the others.
Holography generally requires light sources with high temporal and spatial coherence (such as lasers) for both recording and reconstruction. Variations like volume holography (where interference fringes are recorded throughout the depth of a thick medium) and rainbow holography (which allows viewing in white light) address some of these limitations.

Generally, we differentiate two types of hologram based on recording geometry:

An example of a transmission hologram setup is shown in the next figure (source).

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/Pasted image 20250331170450.png|700
Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/Pasted image 20250331170456.png|700

In ordinary photography, only the intensity distribution |Uo|2 is recorded, so all phase information is lost, and thus three-dimensional reconstruction is not possible.


5.8 Paraxial Ray Optics

Often, for analysing simple optical systems, a full wave-optical (Fourier) treatment is not necessary, and the simpler ray approximation (geometric optics) suffices. This is especially true when effects of diffraction can be neglected, for instance, when all apertures and beam sizes are much larger than the wavelength of light. We define a ray as the local normal to a wavefront. We will work within the paraxial approximation, meaning all rays form small angles with respect to the optical axis (conventionally the z-axis).
A single ray at a transverse plane can be described by a 2D vector, commonly its radial distance r from the axis and its angle θ with respect to that axis (or dr/dz). We will assume systems with cylindrical symmetry around the z-axis for this ray description. For example, a thin converging lens with focal length f transforms an incident ray (r1,θ1) to an output ray (r2,θ2) according to:

r2=r1,θ2=θ1r1/f,

where we have used the small angle approximation tanθsinθθ.

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/QE_script 19.webp|700

This transformation can be expressed using a matrix, called the ABCD matrix or ray-transfer matrix, which relates the output ray vector to the input ray vector:

(r2θ2)=(ABCD)(r1θ1).

For the thin lens described above, the ray-transfer matrix is:

Mlens=(101/f1).

Similarly, for propagation through a homogeneous medium of length d, a ray (r1,θ1) becomes (r2,θ2) where r2=r1+dθ1 and θ2=θ1. The ray-transfer matrix is:

Mprop=(1d01).

For a planar interface between homogeneous media with refractive indices n1 and n2 (ray incident from n1 to n2 at near-normal incidence):

Minterface=(100n1/n2).

The effect of a spherical mirror with concave radius of curvature R (light incident from left, vertex at origin, centre of curvature at z=R) is equivalent to a lens with f=R/2:

Mmirror=(102/R1).

The advantage of this matrix formalism is that the overall ray-transfer matrix for a cascade of optical elements is found by multiplying the individual matrices in the correct order (last element encountered by the ray first in the matrix product if input is on the right, or first element first if input is on the left, depending on how the output vector is written). For example, consider propagation through a medium of length a, followed by a lens of focal length f, then another propagation through length b:

Attachments/Notes/2025 Quantum Electronics/5 Fourier Optics/02_Fundamental_phenomena 24.webp|700

The matrix Mtotal relating the output ray (r,θ) to the input ray (r,θ) is:

(rθ)=Mprop(b)Mlens(f)Mprop(a)(rθ)=(1b01)(101/f1)(1a01)(rθ).