Angular resolution

Source: Wikipedia, the free encyclopedia.
microarcseconds
, comparable to viewing a tennis ball on the Moon (magnification from top left corner counter−clockwise to the top right corner).

Angular resolution describes the ability of any

image-forming device such as an optical or radio telescope, a microscope, a camera, or an eye, to distinguish small details of an object, thereby making it a major determinant of image resolution. It is used in optics applied to light waves, in antenna theory applied to radio waves, and in acoustics applied to sound waves. The colloquial use of the term "resolution" sometimes causes confusion; when an optical system is said to have a high resolution or high angular resolution, it means that the perceived distance, or actual angular distance, between resolved neighboring objects is small. The value that quantifies this property, θ, which is given by the Rayleigh criterion, is low for a system with a high resolution. The closely related term spatial resolution refers to the precision of a measurement with respect to space, which is directly connected to angular resolution in imaging instruments. The Rayleigh criterion shows that the minimum angular spread that can be resolved by an image forming system is limited by diffraction to the ratio of the wavelength of the waves to the aperture width. For this reason, high resolution imaging systems such as astronomical telescopes, long distance telephoto camera lenses and radio telescopes
have large apertures.

Definition of terms

Resolving power is the ability of an imaging device to separate (i.e., to see as distinct) points of an object that are located at a small

angular distance or it is the power of an optical instrument to separate far away objects, that are close together, into individual images. The term resolution or minimum resolvable distance is the minimum distance between distinguishable objects in an image, although the term is loosely used by many users of microscopes and telescopes to describe resolving power. As explained below, diffraction-limited resolution is defined by the Rayleigh criterion as the angular separation of two point sources when the maximum of each source lies in the first minimum of the diffraction pattern (Airy disk) of the other. In scientific analysis, in general, the term "resolution" is used to describe the precision
with which any instrument measures and records (in an image or spectrum) any variable in the specimen or sample under study.

The Rayleigh criterion

Airy diffraction patterns generated by light from two point sources passing through a circular aperture, such as the pupil of the eye. Points far apart (top) or meeting the Rayleigh criterion (middle) can be distinguished. Points closer than the Rayleigh criterion (bottom) are difficult to distinguish.

The imaging system's resolution can be limited either by

Airy pattern, if the wavefront
of the transmitted light is taken to be spherical or plane over the exit aperture.

The interplay between diffraction and aberration can be characterised by the

Lord Rayleigh: two point sources are regarded as just resolved when the principal diffraction maximum (center) of the Airy disk of one image coincides with the first minimum of the Airy disk of the other,[1][2] as shown in the accompanying photos. (In the bottom photo on the right that shows the Rayleigh criterion limit, the central maximum of one point source might look as though it lies outside the first minimum of the other, but examination with a ruler verifies that the two do intersect.) If the distance is greater, the two points are well resolved and if it is smaller, they are regarded as not resolved. Rayleigh defended this criterion on sources of equal strength.[2]

Considering diffraction through a circular aperture, this translates into:

where θ is the angular resolution (

Bessel function of the first kind
divided by
π.

The formal Rayleigh criterion is close to the

image processing techniques including deconvolution
of the point spread function allow resolution of binaries with even less angular separation.

Using a small-angle approximation, the angular resolution may be converted into a spatial resolution, Δ, by multiplication of the angle (in radians) with the distance to the object. For a microscope, that distance is close to the focal length f of the objective. For this case, the Rayleigh criterion reads:

.

This is the

Fourier properties
of a lens.

A similar result holds for a small sensor imaging a subject at infinity: The angular resolution can be converted to a spatial resolution on the sensor by using f as the distance to the image sensor; this relates the spatial resolution of the image to the f-number, f/#:

.

Since this is the radius of the Airy disk, the resolution is better estimated by the diameter,

Specific cases

Log–log plot of aperture diameter vs angular resolution at the diffraction limit for various light wavelengths compared with various astronomical instruments. For example, the blue star shows that the Hubble Space Telescope is almost diffraction-limited in the visible spectrum at 0.1 arcsecs, whereas the red circle shows that the human eye should have a resolving power of 20 arcsecs in theory, though normally only 60 arcsecs.

Single telescope

Point-like sources separated by an

arcsecond, but astronomical seeing
and other atmospheric effects make attaining this very hard.

The angular resolution R of a telescope can usually be approximated by

where λ is the

nm
, for a resolution of 0.1 arc second, we need D=1.2 m. Sources larger than the angular resolution are called extended sources or diffuse sources, and smaller sources are called point sources.

This formula, for light with a wavelength of about 562 nm, is also called the Dawes' limit.

Telescope array

The highest angular resolutions for telescopes can be achieved by arrays of telescopes called astronomical interferometers: These instruments can achieve angular resolutions of 0.001 arcsecond at optical wavelengths, and much higher resolutions at x-ray wavelengths. In order to perform aperture synthesis imaging, a large number of telescopes are required laid out in a 2-dimensional arrangement with a dimensional precision better than a fraction (0.25x) of the required image resolution.

The angular resolution R of an interferometer array can usually be approximated by

where λ is the

baseline. The resulting R is in radians
. Sources larger than the angular resolution are called extended sources or diffuse sources, and smaller sources are called point sources.

For example, in order to form an image in yellow light with a wavelength of 580 nm, for a resolution of 1 milli-arcsecond, we need telescopes laid out in an array that is 120 m × 120 m with a dimensional precision better than 145 nm.

Microscope

The resolution R (here measured as a distance, not to be confused with the angular resolution of a previous subsection) depends on the angular aperture :[5]

where .

Here NA is the numerical aperture, is half the included angle of the lens, which depends on the diameter of the lens and its focal length, is the refractive index of the medium between the lens and the specimen, and is the wavelength of light illuminating or emanating from (in the case of fluorescence microscopy) the sample.

It follows that the NAs of both the objective and the condenser should be as high as possible for maximum resolution. In the case that both NAs are the same, the equation may be reduced to:

The practical limit for is about 70°. In a dry objective or condenser, this gives a maximum NA of 0.95. In a high-resolution

nm. Given that the shortest wavelength of visible light is violet
(),

which is near 200 nm.

Oil immersion objectives can have practical difficulties due to their shallow depth of field and extremely short working distance, which calls for the use of very thin (0.17 mm) cover slips, or, in an inverted microscope, thin glass-bottomed Petri dishes.

However, resolution below this theoretical limit can be achieved using super-resolution microscopy. These include optical near-fields (Near-field scanning optical microscope) or a diffraction technique called 4Pi STED microscopy. Objects as small as 30 nm have been resolved with both techniques.[6][7] In addition to this Photoactivated localization microscopy can resolve structures of that size, but is also able to give information in z-direction (3D).

List of telescopes and arrays by angular resolution

Name Image Angular resolution (
arc seconds
)
Wavelength Type Site Year
Global mm-VLBI Array (successor to the Coordinated Millimeter VLBI Array) 0.000012 (12 μas) radio (at 1.3 cm)
very long baseline interferometry array of different radio telescopes
a range of locations on Earth and in space[8] 2002 - 
PIONIER
0.001 (1 mas) light (1-2 micrometre)[9] largest
optical array of 4 reflecting telescopes
Paranal Observatory, Antofagasta Region, Chile 2002/2010 -
Hubble Space Telescope 0.04 light (near 500 nm)[10] space telescope Earth orbit 1990 -
James Webb Space Telescope 0.1[11] infrared (at 2000 nm)[12] space telescope Sun–Earth L2 2022 -

See also

Notes

  1. ^ In the case of laser beams, a Gaussian Optics analysis is more appropriate than the Rayleigh criterion, and may reveal a smaller diffraction-limited spot size than that indicated by the formula above.

References

  1. ^ .
  2. ^ a b Lord Rayleigh, F.R.S. (1879). "Investigations in optics, with special reference to the spectroscope". .
  3. ^ Michalet, X. (2006). "Using photon statistics to boost microscopy resolution".
    PMID 16549771
    .
  4. ^ "Diffraction: Fraunhofer Diffraction at a Circular Aperture" (PDF). Melles Griot Optics Guide. Melles Griot. 2002. Archived from the original (PDF) on 2011-07-08. Retrieved 2011-07-04.
  5. ^ Davidson, M. W. "Resolution". Nikon’s MicroscopyU. Nikon. Retrieved 2017-02-01.
  6. ^ Pohl, D. W.; Denk, W.; Lanz, M. (1984). "Optical stethoscopy: Image recording with resolution λ/20". .
  7. ^ Dyba, M. "4Pi-STED-Microscopy..." Max Planck Society, Department of NanoBiophotonics. Retrieved 2017-02-01.
  8. ^ "Images at the Highest Angular Resolution in Astronomy". Max Planck Institute for Radio Astronomy. 2022-05-13. Retrieved 2022-09-26.
  9. .
  10. ^ "Hubble Space Telescope". NASA. 2007-04-09. Retrieved 2022-09-27.
  11. ].
  12. ^ "FAQ Full General Public Webb Telescope/NASA". jwst.nasa.gov. 2002-09-10. Retrieved 2022-09-27.

External links